54
Review Isolation of Industrial Important Bioactive Compounds from Microalgae Vimala Balasubramaniam 1 , Rathi Devi-Nair Gunasegavan 1 , Suraiami Mustar 1 , Lee June Chelyn 2 and Mohd Fairulnizal Md Noh 1 1 Nutrition, Metabolism & Cardiovascular Research Centre, Institute for Medical Research, NIH, Level 3, Block C7, No.1, Jalan Setia Murni U13, Setia Alam, 40170 Shah Alam, Selangor Malaysia 2 Herbal Medicine Research Centre, Institute for Medical Research, Institute for Medical Research, NIH, Level 3, Block C7, No.1, Jalan Setia Murni U13, Setia Alam, 40170 Shah Alam, Selangor Malaysia * Correspondence: [email protected]; [email protected] Abstract: Microalgae are known as a rich source of bioactive compounds which exhibit different biological activities. Increased demand for sustainable biomass for production of important bioactive components with various potential especially therapeutic applications has resulted in noticeable interest in algae. Utilisation of microalgae in multiple scopes has been growing in various industries ranging from harnessing renewable energy to exploitation of high-value products. The focuses of this review are on production and use of value-added components obtained from microalgae with current and potential application in the pharmaceutical, nutraceutical, cosmeceutical, energy and agri-food industries, as well as for bioremediation. Moreover, this work discusses the advantage, potential new beneficial strains, applications, limitations, research gaps and future prospect of microalgae in industry. Keywords: microalgae; industry; isolation; bioactive compounds; nutraceuticals; pharmaceutical; cosmeceutical 1. INTRODUCTION Microalgae are in the form of unicellular, multicellular, filamentous or siphonaceous, known as photosynthetic microorganism that can be categorized as eukaryotic and prokaryotic [1]. Microalgae are also the largest global primary producers consist of approximately 200,000 species [2] with distinctive nutrient contents as well as bioactive compounds which have a wide spectrum of commercial applications in various facets of industries including pharmaceuticals, nutraceuticals, cosmeceuticals, biofuels, biofertilisers, wastewater treatments, feed, and proteomics (Figure 1). Production of microalgae involves mass cultivation, recovery of biomass and downstream processes for sustainable yield to cater for food, chemical, feed, biofuel, and high value products. Intrinsic factors such as temperature, salinity, light and the availability of nutrients affect the chemical composition of the biomasses. Figure 2, illustrates a few of the cultivation processes of microalgae. The applications of microalgae in industries concentrated to a few specific species which has high economic value. The highly-sought genera in the global algae market were dominated by Spirulina and Chlorella in the form of dried biomass due to various beneficial health effects [3]. Among the myriad components which were exploited for commercial purpose are fatty acids, carotenoids, vitamin, minerals, polysaccharides and bioactive compounds. According to the latest analysis, the global market for microalgae is forecasted to reach USD 3,318 million by 2022 driven mainly by the demand from pharmaceutical and nutraceutical industries [4] owing to customers’ increasing health concern, interest on natural alternatives as well as escalating chronic diseases. Likewise, various microalgae derived compounds reported to exert various skin benefits which currently gaining attention in many aspects of cosmeceutical production. Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1 © 2020 by the author(s). Distributed under a Creative Commons CC BY license.

Review Isolation of Industrial Important Bioactive

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Review Isolation of Industrial Important Bioactive

Review

Isolation of Industrial Important Bioactive

Compounds from Microalgae

Vimala Balasubramaniam 1, Rathi Devi-Nair Gunasegavan 1, Suraiami Mustar 1, Lee June Chelyn 2 and Mohd Fairulnizal Md Noh 1

1 Nutrition, Metabolism & Cardiovascular Research Centre, Institute for Medical Research, NIH, Level 3,

Block C7, No.1, Jalan Setia Murni U13, Setia Alam, 40170 Shah Alam, Selangor Malaysia 2 Herbal Medicine Research Centre, Institute for Medical Research, Institute for Medical Research, NIH,

Level 3, Block C7, No.1, Jalan Setia Murni U13, Setia Alam, 40170 Shah Alam, Selangor Malaysia

* Correspondence: [email protected]; [email protected]

Abstract: Microalgae are known as a rich source of bioactive compounds which exhibit different

biological activities. Increased demand for sustainable biomass for production of important

bioactive components with various potential especially therapeutic applications has resulted in

noticeable interest in algae. Utilisation of microalgae in multiple scopes has been growing in various

industries ranging from harnessing renewable energy to exploitation of high-value products. The

focuses of this review are on production and use of value-added components obtained from

microalgae with current and potential application in the pharmaceutical, nutraceutical,

cosmeceutical, energy and agri-food industries, as well as for bioremediation. Moreover, this work

discusses the advantage, potential new beneficial strains, applications, limitations, research gaps

and future prospect of microalgae in industry.

Keywords: microalgae; industry; isolation; bioactive compounds; nutraceuticals; pharmaceutical;

cosmeceutical

1. INTRODUCTION

Microalgae are in the form of unicellular, multicellular, filamentous or siphonaceous, known as

photosynthetic microorganism that can be categorized as eukaryotic and prokaryotic [1]. Microalgae

are also the largest global primary producers consist of approximately 200,000 species [2] with

distinctive nutrient contents as well as bioactive compounds which have a wide spectrum of

commercial applications in various facets of industries including pharmaceuticals, nutraceuticals,

cosmeceuticals, biofuels, biofertilisers, wastewater treatments, feed, and proteomics (Figure 1).

Production of microalgae involves mass cultivation, recovery of biomass and downstream processes

for sustainable yield to cater for food, chemical, feed, biofuel, and high value products. Intrinsic

factors such as temperature, salinity, light and the availability of nutrients affect the chemical

composition of the biomasses. Figure 2, illustrates a few of the cultivation processes of microalgae.

The applications of microalgae in industries concentrated to a few specific species which has

high economic value. The highly-sought genera in the global algae market were dominated by

Spirulina and Chlorella in the form of dried biomass due to various beneficial health effects [3]. Among

the myriad components which were exploited for commercial purpose are fatty acids, carotenoids,

vitamin, minerals, polysaccharides and bioactive compounds. According to the latest analysis, the

global market for microalgae is forecasted to reach USD 3,318 million by 2022 driven mainly by the

demand from pharmaceutical and nutraceutical industries [4] owing to customers’ increasing health

concern, interest on natural alternatives as well as escalating chronic diseases. Likewise, various

microalgae derived compounds reported to exert various skin benefits which currently gaining

attention in many aspects of cosmeceutical production.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

© 2020 by the author(s). Distributed under a Creative Commons CC BY license.

Page 2: Review Isolation of Industrial Important Bioactive

2

1

Microalgae Biomass

Pharmaceutical

Anticancer

Antiinflammatory

Antiviral

Antifungal

Antibacteria

Neutraceutical/

Food

Food

supplement

Gelling

agent

Food

additives

Bio-colorant

Cosmeceutical

Skin

whitening

Skin

disease

Antioxidant

Moisturising agent

Fuel/Energy

Biodiesel

Bioethanol

Electricity

/Heat

Biogas

Fertilizer

Enhance soil fertility

Plant growth

stimulant

Bio-pesticide

ProteomicsWaste water

treatmentFeed

Nutritional source

Enhance fertility

Enhance immune system

Figure 1. Potential uses of microalgae in various industries.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 3: Review Isolation of Industrial Important Bioactive

3

Figure 2. Lab scale microalgae cultivation types. A) Raceway photobioreactor B) Indoor annular

column photobioreactor, C) Microalgae culture in bottles with light and oxygen in Scientific

laboratory, D) Open annular photobioreactor, (Courtesy: Dr. Adibi Rahiman Md Nor, University

Malaya).

Advances of new application areas of microalgae in aquaculture and biofuel production have

provided a significant rise of algae demand in the global market. Microalgae are constituted of high

level of lipids. Various research has highlighted the potential of microalgae biomass as a source of

renewable energy, namely biofuel which is imperative to reduce the dependency on fossil fuel [5].

Microalgae have an upper hand in the biofuel production compared to other bioenergy sources (corn,

sugar cane, palm oil and etc.) as they do not require arable land for cultivation, thus eliminating

competition for space and resources with food crops. Some of the key players in the algae industries

were Algae Tec, Pond Biofuels Incorporated, Cyanotech, Kai BioEnergy, Algae Systems and others

[4].

The review aims to summarise the value-added components from microalgae with potential

application in the pharmaceutical, nutraceutical, cosmeceutical, energy and agri-food industries, as

well as for bioremediation along with commercial applications and examples of microalgal

manufacturers as well as commercialized products.

2. MICROALGAE BIOMASS

The mass production of algae biomass is important for various industries [6]. Numerous

methods have been established for microalgae-based products development and down-stream

processes with the advancement in the technologies in this area.

2.1. Biomass Production

A B

B

B

C D

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 4: Review Isolation of Industrial Important Bioactive

4

The process of biomass production of microalgae encloses several steps such as cultivation,

harvesting and biomass dehydration as shown in Figure 3

Figure 3. The production of microalgae biomass [5,6]

2.1.1 Cultivation

There are two systems developed for the production or culturing of algal biomass; the open

pond and closed photobioreactor (PBR) technologies. Open pond production is categorised into two

systems; natural waters (ponds, lakes and lagoons) [7] and artificial ponds (circular and raceway)

[7,8]. The open pond is a cheaper method of large-scale algal biomass production compared to the

PBR. The PBR however, provide an excellent and controlled closed culture systems for cultivation,

preventing hazard or contamination from moulds, bacteria, protozoa and competition by other

microalgae [9]. It is usually placed outdoor to exploit the free sources of energy from sunlight. There

are three types of PBR categorised into tubular (TPBR), vertical column (VCPBR) and flat-plate (FP-

PBR) [10].

2.1.2 Harvesting

The microalga biomass can be separated from the culture medium or harvested by four means;

biomass aggregation (flocculation and ultrasound), flotation, centrifugation and filtration. In some

cases, combinations of two or more techniques are used to increase effectiveness. Harvesting method

selection depends on several criteria of the microalgae such as the density, size and the desired final

products [11].

(a) Biomass aggregation: In the flocculation technique, microalgae cells are aggregated together to

form a larger particle known as floc, with the addition of flocculants such as multivalent cations and

cationic polymers to the media which help neutralised the cells surface charge [12]. There are two

Cultivation

• Open pond (Natural and Artificial)

• Closed photobioreactor, PBR (Tubular, Vertical column and Flat-plate)

Harvesting

• Biomass aggregation (Flocculation and Ultrasound)

• Flotation

• Centrifugation

• Filtration

Dehydration

• Sun-drying

• Spray-drying

• Freeze-drying or Lyophilisation

MICROALGAE BIOMASS PRODUCTION

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 5: Review Isolation of Industrial Important Bioactive

5

types of flocculating agents; chemical and bio-flocculants. The cheaper and easily available chemical

flocculants that are widely used in industry are iron and aluminium salts [13]. However, toxic effects

can occur when using chemical, causing contamination to the biomass. The bio-flocculants which are

safer, eco-friendly and no pre-treatment needed for further downstream processing are the most-

preferred [19,20]. The criteria used to choose flocculants include nontoxic, low-cost and effective at

low concentration. Among Meanwhile, the common biopolymer bio-flocculants used include acrylic

acid and chitosan [14]. In the ultrasound technique, instigates aggregation is initiated followed by

increased sedimentation to facilitate harvesting of the algal biomass [15]. The benefit of using this

method is that the valuable metabolites are preserved because ultrasonic harvesting does not produce

shear stress on the biomass although used continuously [16].

(b) Flotation: It is a technique intended to float algal cells on the surface of the water, using a micro-

air bubbles disperser without the addition of chemicals [17]. This technique is economically

advantage due to the low operational costs with an easy operating procedure and high harvesting of

biomass [18].

(c) Centrifugation: Centrifugation It is the recovery of algal biomass from the culture media using

centrifuge by gravitational force [19]. This technique is rapid, easy and efficient but the cost can

escalate due to the high energy input and maintenance required [20]. Another disadvantage of this

technique is the internal damage of cells, causing loss of delicate nutrients if a high gravitational force

is used [21].

(d) Filtration: Filtration It is a process to isolate alga biomass from the liquid culture medium by

using a porous membrane with various particle size range [22]. It can be implemented through three

different ways; conventional, microfiltration and ultrafiltration (isolation of metabolites). The

conventional filtration is used to harvest large size microalgae (> 70μm) such as Coelastrum and

Spirulina. Microfiltration and ultrafiltration are used to harvest smaller size microalgae, equal to the

size of the bacteria [23]. Filtration may be relatively slow compared to using centrifugation. However,

processing smaller media volumes require lower cost than centrifugation. But in a large-scales

production, centrifugation can be more economical compared to using filtration due to the cost of

membrane replacement and pumping [30].

2.1.3 Biomass Dehydration

Algae biomass is immediately processed to the following stage after being separated from the

culture medium to prevent spoilage or to extend their shelf-life [22]. Three different types of drying

or dehydration process that are normally used include sun-drying, spray-drying and freeze-drying.

The method chosen is entirely dependent on the desired final products.

(a) Sun-drying: It is the cheapest method available compared to the other two techniques. This

technique is solely based on the solar energy which causes limitations in terms of weather condition,

long drying period and the large drying area needed [24]. Since drying using sunlight is an

uncontrollable process, the problem of overheating may occur, change of texture, colour and taste of

the microalgae [22].

(b) Spray-drying: This technique is to produce dry powder from a fine spray of suspension droplets

which is in continuous contact with hot air in a large vessel. This method has many advantages such

as can be operated continuously, the powder produced is very fine and the rapid drying can maintain

a good quality product [25,26]. This method is usually opted for high-value operations due to its

efficiency, but some algal components such as pigments can be significantly deteriorate and the

operation cost is expensive [22].

(c)Freeze-drying or Lyophilisation: It is widely used at laboratory-scale only to dry microalgae since

large scale production can be very expensive [20]. Freeze-drying is a direct dehydration process of

frozen products using sublimation mechanism. The microalgae are frozen to solidify the material

within before freeze-drying. The moisture content of the microalgae is decreased slowly at low-

temperature, maintaining the solid structure and the quality of the product [27].

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 6: Review Isolation of Industrial Important Bioactive

6

2.2 Extraction of Bioactive Compound

The microalgae are composed of carbohydrates, lipids, proteins, minerals and many other

compounds. To utilise the compounds for different application such as for biofuels/energy and

agricultural use, the algal biomass will be pre-treated to release the stored bioactive compound in the

cells [28]. The cell walls will be lysed to enable all the desired components to be extracted and this

may be accomplished by various ways, such as physical, mechanical (bead milling, homogenisation,

microwave, ultrasonic and pulsed electric field), chemical (solvent, acid and alkali) and biological

(enzymes) methods. The method chosen for the pre-treatment process is based on the desired final

products [29,30].

3. MICROALGAE IN PHARMACEUTICAL

Microalgae are potential source for bioactive components with pharmaceutical applications. Several

important microalgae-derived components with their pharmaceutical applications are highlighted

in Table 1.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 7: Review Isolation of Industrial Important Bioactive

7

Table 1. Important micro-algae derived components in pharmaceutical applications

Compound Microalgae species Isolation/

Extraction method

Cell/ virus/animal model/ clinical

patient

Effect References

Anticancer agents

Lutein Chlorella vulgaris Ethanol extraction and

partitioned with hexane

Colon cancer (HCT116) cells Antiproliferative effect

IC50=40.31±4.43 µg/mL

[31]

Violaxanthin Dunaliella tertiolecta Dichloromethane extract Human mammary carcinoma cell lines

(MCF-7) and human prostatic carcinoma (LNCaP) cells

Potent inhibition of MCF-7 and

LNCaP cells at growth inhibition (GI50) of 56.1 and 60.9 µg/mL,

respectively

[32]

Fucoxanthin Phaeodactylum tricornutum UTEX 640

Pressurized liquid extraction

Human liver cancer cell lines (Hep-G2), Human colon cancer (Caco-2) cells and

HeLa cell line

An inhibitory effect of up to 58% was measured in Hep-G2 cells.

In HeLa and Caco-2 cells, the effect was stronger than that of the

positive control with a final

concentration of 5% DMSO.

[33]

Phycocyanin Spirulina platensis Freeze-thawing followed

by solvent extraction

Liver cancer (Hep-G2) cancer cells, Non-

small cell lung cancer (NSCLC)

Inhibited liver cancer cells and

leukemia cells.

Induce apoptosis in H460 cells

reaching 3.72±0.98% and suppress

growth of NSCLC cells

[34-37]

Phycocyanin Spirulina platensis Supercritical fluid extraction

Human lung cancer cells (A549) IC50=26.82 µg/mL [38]

Phycocyanin Limnothrix sp

37-2-1

Fractional precipitation and

purification with activated charcoal and chitosan

Prostate Cell Line LNCaP C-PC alone (250 and 500 µg/mL)

killed 65% and 70% of cells while C-PC (500 µg/mL) combined with

Topetencan (TPT) (1 µM) killed 80% of cells due to additive effect

[39]

Phycocyanin Limnothrix sp.

NS01

Four-step purification procedure including the

adsorption of impurities

with chitosan, activated charcoal, ammonium

sulfate precipitation, and

ion-exchange chromatography

Human breast cancer cell line (MCF-7) IC50 for 24, 48 and 72 hr exposure to C-PC were 5.92, 5.66 and 4.52

µg/µL

[40]

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 8: Review Isolation of Industrial Important Bioactive

8

Table 1. ( Continued)

Compound Microalgae species Isolation/

Extraction method

Cell/ virus/animal model/ clinical

patient

Effect References

Cardioprotective agents

cis β-carotene

Dunaliella bardawil

(containing a mixture of cis and trans-isomers)

Enriched extract Mice fed a high-fat diet In old mice with established

atherosclerotic lesion, Dunaliella

inhibited significantly plasma cholesterol elevation and atherosclerosis progression

[41]

Eicosapentanoic acid (EPA) Nannochloropsis Enriched EPA oil Healthy subjects Lower cholesterol levels [42]

Antiviral agent

Cyanovirin Nostoc ellpsosporum Aqueous extraction HIV-1 laboratory strains;

HIV-1 primary isolates;

HIV-2; SIV;

HIV-1 laboratory strains (EC50 0.1-

5.8 nM), HIV-1 primary isolates

(EC50 1.5-36.8 nM), HIV-2 (EC50 2.3-7.6 nM), SIV (EC50 11 nM)

[43]

Scytovirin Scytonema varium Aqueous extraction HIV-1 laboratory strain (HIV-1RF); HIV-1 primary isolates (ROJO) in

peripheral blood mononuclear cell

(PBMC); HIV-1 primary isolates (Ba-L and ADA)

in macrophages

HIV-1RF (EC50 0.3 nM), HIV-1 primary isolates (ROJO) in

PBMC (EC50 7 nM), HIV-1 primary

isolates (Ba-L) (EC50 22 nM), HIV-1 primary isolates (ADA) (EC50 17

nM)

[44]

Mixture of Icthypeptins A and Icthypeptins

Microcystic ichthyoblabe Methanol extract Influenza A IC50 12.5 µg/mL [45]

Calcium spirulan Spirulina platensis Aqueous extraction HIV-1;

Herpes simplex virus type 1 (HSV-1)

HIV-1 (IC50 9.3 µg/mL),

Herpes simplex virus type 1 (HSV-1) (IC50 9.0 µg/mL).

[46]

Sulfated exopolysaccharide Porphyridium cruentum Aqueous extraction Herpes simplex virus type 1 (HSV-1); Herpes simplex virus type 2 (HSV-2);

Vericella virus (VZV)

HSV-1 (CPE50 protection 1 µg/mL), HSV-2 (CPE50 protection 5 µg/mL), VZV (CPE50 protection 0.7 µg/mL).

[47]

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 9: Review Isolation of Industrial Important Bioactive

9

Many studies have documented the health benefits of microalgae compounds for prevention and improvement of diseases such as diabetes, obesity,

cardiovascular disease, cancer, inflammation, Alzheimer’s diseases, depression as well as bacterial, fungal and viral infections [42,48,49]. Despite that, only a

limited number of microalgae with pharmaceutical applications are currently available as listed in Table 2. Low extraction yield and high production cost are some

of the factors in delaying commercialization of some microalgae-derived bioactives [50]. In addition, there are potential risk of severe side effects, allergic

reactions and accumulation of heavy metals and toxins in some species of microalgae [51]. Consequently, strong emphasis on the good manufacturing

practices in cultivation, harvesting, extracting and purification and controls to limit toxin and impurities are required to ensure safety, efficacy and quality of

microalgae-derived purified compounds and enriched extracts for approval and commercialization [52].

Table 2. Examples of bioactive components from microalgae that are produced in commercial scale

Microalgae Bioactive Product Pharmaceutical

applications

References

Haematococcus pluvialis

Astaxanthin

Spirulina (Earth Spirulina Group, ES Co, Seoul, Korea)

Lipid lowering [53]

Schizochytrium limacinum Docosahexaenoic acid (DHA) Maris DHA oil (IOI, Hamburg,Germany) Rheumatoid arthritis [54]

Nannochloropsis Eicosapentaenoic acid (EPA)

Almega®PL (iWi Life) (Qualitas Health, Houston, US)

Cholesterol lowering

[42]

Arthrospira FEM-101 Allophycocyanin ApoX surface antiviral spray (FEBICO, Taiwan) Antiviral [55]

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 10: Review Isolation of Industrial Important Bioactive

10

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 11: Review Isolation of Industrial Important Bioactive

11

3.1 Compounds with anti-cancer properties

Microalgae-derived bioactives with anti-cancer properties are extensively studied in recent years

[33]. Among the compounds, the microalgae pigments such as astaxanthin, β-carotene, lutein,

violaxanthin and fucoxanthin have most potential to be commercialized as pharmaceuticals because

of their established applications as nutraceuticals and cosmetics and the rising demand as dietary

supplements. In fact, astaxanthin and β-carotene are currently produced commercially from

microalgae Dunaliella salina (D. salina) and Hamatococcus pluvialis, respectively while commercial

production of other carotenoids such as lutein and fucoxanthin are gaining momentum [56].

Fucoxanthin for instance, have been isolated from diatom microalgae Phaeodactylum tricornutum

which is cultivated in a pilot-scale photobioreactor and can be considered as a commercially viable

source for fucoxanthin [51]. The anti-cancer properties of some of these carotenoids are summarized

in Table 1. Extraction of carotenoid is achieved using ultra-sound extraction and freeze-thawing

methods however organic solvent extraction at high temperature and pressure are more widely used

method in commercial- scale production [56].

Another promising anticancer agent is phycocyanin, a protein pigment from phycobiliprotein

group. Phycocyanins are isolated from the commercially grown microalgae Spirulina platensis but

isolation from other cyanobacteria such as Limnothrix sp. has also been reported (Table 1) [41,43]. C-

phycocyanin showed inhibitory activity in liver cancer cell lines (HepG2) [57], human leukemia cells

(K562) [58] and against lung cancer cell lines (A549 and NSCLC) [39,40,41]. In another study, the

phycocyanin from Limnothrix sp. enhanced the anticancer properties of the anticancer drug Topetecan

against prostate cancer cell line (LNCap) [42]. Meanwhile, phycocyanin isolated from as Limnothrix

sp. NS01 with 2 subunit α and β (17 and 20 kDa) showed antiproliferative activity in human breast

cancer cell lines (MCF-7) [43]. Isolation of C-phycocyanin from Spirulina platensis was accomplished

using buffer ammonium sulfate solution (i.e. salting- out technique) as well as by using supercritical

fluid extraction using ethanol as a modifier which resulted in higher yield compared to conventional

solvent extraction [41,57]. In the studies, impurities were removed using chitosan and activated

charcoal however further purified the compound by ion-exchange chromatography was also carried

out [43].

3.2 Compounds with cardioprotective properties

Several microalgae-derived compounds have also been studied for its cardioprotective effect

and briefly summarized in Table 1. Carotenoids are shown to possess antioxidant properties that are

important for preventing cell damage caused by free radicals associated with chronic cardiovascular

diseases and stroke [59]. In this regard, some microalgae producing high amount of carotenoids have

been investigated for their cardioprotective properties. For instance, D. salina microalgae can produce

up to 10-13% of β-carotene have been shown to have protective effect against atherosclerosis in both

mouse and humans. Furthermore, a mixture of tran-isomers (~40%) and cis β-carotene isomers (~60%)

from D. salina was found to be more potent in decreasing total lipid, cholesterol and triglyceride (TG)

levels compared to all trans-β-carotene found in synthetic β-carotene [60]. Harari et al. [44] also

showed that cis β-carotene isomers (~50%) from Dunaliella bardawil powder inhibited atherosclerosis

progression in older mice with a high-fat diet. Although cis β-carotene is still not produced

commercially due to its high production costs, a high density inoculum enriched with cis β-carotene

strain from D. salina are currently being developed to deliver reproducible, low-cost D. salina biomass

containing high content of 9-cis β-carotene [31].

Another group of bioactive with cardioprotective properties are the polyunsaturated fatty acids

(PUFAs) especially the omega-3 fatty acids such as DHA, EPA and α-linoleic acid (ALA) which have

been shown to reduce blood cholesterol and improve hypertension. Of these fatty acids, DHA is the

only PUFA currently commercially available. Purified EPA sourced from various microalgae

including Porphyridium purpureum and Isochrysis galbana are still not economically competitive to be

produced commercially. However, a proprietary strain of microalgae Nannochloropsis cultivated in

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 12: Review Isolation of Industrial Important Bioactive

12

an open pond with high solar radiation was shown to produce high EPA content (>65%) oil marketed

as A2 EPA PureTM for supplement and pharmaceutical applications have been reported [61]. PUFAs

are commercially extracted using hexane solvent followed by mechanical pressing. However,

extracted PUFAs are prone to oxidation and therefore all materials that can initiate oxidation such as

copper are eliminated from the extraction and storage area [61].

3.3 Compounds with antiviral properties

Microalgae are also potential sources for bioactives with antiviral properties (Table2). For

example, lectin protein with antiviral properties such as cyanovirin (CV-N) and scytovirin (SVN)

have been reported [46,47]. CV-N is a 11kDa protein consisting of 101-amino acid in single chain with

two disulfide linkages and was isolated from the aqueous extract of Nostoc ellpsosporum and purified

by ethanol precipitation followed by fractionation and purification by column chromatography [62].

CV-N showed a broad spectrum of antiviral activity against human immunodeficiency virus-1 (HIV

type-1) laboratory and clinical strains with effective concentrations (EC50) range 0.1-5.8 nM and 1.5-

36.8 nM, respectively [46]. In the same study, CV-N was also active against human immunodeficiency

virus-2 (HIV-2 ) and simian immunodeficiency virus with EC50 2.3-7.6 nM and 11 nM, respectively

[46]. Although CV-N showed promising antiviral activity including in an animal HIV transmission

model the clinical application of this molecule is currently limited due to its reported mitogenic

activity [63]. Meanwhile, another lectin protein SVN has been isolated from the aqueous extracts of

cultured cyanobacterium Scytonema varium. SVN is a 9.71kDa protein consisting of 95-amino acid

chains with 5-disulfide linkage which also showed potent antiviral activity against HIV type-1

laboratory strains and clinical isolates with EC50 between 0.3-22 nM [47].

Another group of compounds with antiviral activities are the cyclic peptides. The fractions

containing a mixture of Icthypeptins A and Icthypeptins B, cyclic depsipeptides isolated from

Microcystic ichthyoblabe showed antiviral activity against influenza A virus with inhibitory

concentration (IC50) of 12.5µg/mL comparable to the control amantadine IC50 of 15µg/mL [64]. Besides

that, sulfated polysaccharides with antiviral properties have also been described. For example,

calcium spirulan isolated from Spirulina platensis showed antiviral activity against HIV-1 with IC50 of

9.3 µg/ml comparable to the dextran sulfate control when assessed by P24 antigen assay [49]. The

sulfated exopolysaccharide (1 µg/mL) from the Spanish strain of Porphyridium cruentum showed

strong inhibitory effect on the cytopathic effect on Herpes simplex virus-1 (HSV-1), Herpes simplex

virus-2 (HSV-2) and Vericella virus (VZV) with CPE50 protection of 0.7-5 µg/mL [50].

4. MICROALGAE IN NUTRACEUTICALS/FOOD

A number of microalgae have been classified as Generally Regarded as Safe (GRAS) and

approved by US Food and Drug Administration (FDA). For the guaranteed safety and valuable

source of nutrient, algae are used widely in industries especially for food and nutraceutical

applications. According to Watanabe [65], microalgae species of cyanobacteria eg. Spirulina,

Aphanizomenon and Nostoc are hugely harvested for the food industry. Dried Aphanizomenon,

contributes approximately 500 tons annually which is dominantly produced in North America at

Upper Klamath Lake, Klamath Falls, Oregon for food supplement industry while Spirulina is widely

cultured and produced in countries like United States, Taiwan, China, India and others with an

estimated output of 3000 tons annually [66]. Myanmar Spirulina factory in Yangon produces tablets,

chips, pasta and liquid extract [3]. Other species like Chlorella sp, Haematococcus sp, Dunaliella sp.

cultivated commercially by various countries for nutraceuticals and food application. List of

commercialized strain, industry application, companies and the countries involved is provided in

Table 3.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 13: Review Isolation of Industrial Important Bioactive

13

Table 3. List of commercialized strain, industry application, companies and the countries involved

Biomass

Extracted compound Isolation/Extraction method Application in industry Company Country References

Spirulina platensis

Algae protein; Biomass

Enzymatic hydrolysis, Physical processes, Chemical extraction,

Ultrasound-assisted extraction,

Pulsed electric field, and Microwave-assisted extraction

Health supplement, Health food, Infant

formula

Saxony-Anhalt; Parry Nutraceuticals;

Japan Spirulina Co., Ltd;

Siam Alga Co., Ltd.

Germany; India;

Japan;

Thailand

[3,67,68]

Spirulina platensis Vitamin 12 Water extraction method Health supplement Myanmar Spirulina Factory Myanmar [68,69]

Chlorella vulgaris

Chorella sp.

Biomass, pigments Microwave-assisted extraction Health food, food

supplement

Nikken Sohonsha Corp.;

Chlorella manufacturing and

Co.; Klötze;Ocean Nutrition

Japan;

Taiwan; Germany;

Canada

[3,68]

Haematococcus pluvialis Astaxanthin Mechanical treatments, chemical

treatments using solvents, pressurized extraction,

ultrasounds and microwaves

Food supplement, bio-

colorant

Algae Health Science;

Cyanotech Corporation; Aquasearch

Algatechnologies

China;

US; Israel

[68,70-73]

Dunaliella salina Beta-carotene Solvent extraction; Microwave-

assisted extraction

Health food, Dietary

supplement, bio-colourant

Cyanotech;

Earthrise Nutritionals;

Nature Beta Technologies

Cognis; Betadene;

Cognis Nutrition and Health

US;

Australia

[68,74]

Spirulina platensis Phycocyanin Microwave-assisted extraction Bio-colourant, Health supplement

Panmol/Madaus;

Yunnan Green A Biological

Project Co., Ltd

Austria;

China

[68]

Ulkenia sp. DHA Solvent extraction and microwave Dietary supplement; Nutrinova German [3]

Schizochytrium sp. DHA Cellular hydrolysis Dietary supplement; Health food

OmegaTech USA [3]

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 14: Review Isolation of Industrial Important Bioactive

14

Table 3. (Continued)

Biomass

Extracted compound Isolation/Extraction

method

Application in industry Company Country References

Crypthecodinium cohnii DHA Microwave Infant formula Martek USA [75]

Nannochloropsis oculata Omega-3 PUFA Solvent extraction; Microwave

Omega-3 supplements Qualitas;

Cleanalgae SL; Astaxa

USA;

Spain; Germany

[76,77]

Porphyridium spp Polysaccharides Ultrafiltration Food additives, Nutrition

InnovalG France [68,78,79]

Euglena gracilis Biomass Flocculation Health food Euglena Japan [80]

Odontella aurita Fatty acids Health supplement InnovalG France [68]

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 15: Review Isolation of Industrial Important Bioactive

15

54.1 Algal Protein

Demand for plant-based nutrient especially protein sources has augmented over the years owing

to growing health concern and shift of millennial preference across the world to nutraceutical

products which are convenient and offers a high value of nutrition. This, in turn, has the

manufacturers and industries of food and beverages as well as nutraceuticals to search for lucrative

sources of protein. Among all, microalgae showed as a promising source of protein combined with

diverse bioactive compounds and essential nutrients. The demand for global algae protein exceeded

USD700 million in 2019 and projected to expand over the years in view of changing lifestyle of

consumers and preference [81].

Among the myriad of algae, Chlorella and Spirulina species are most sought-after in the global

microalgae market owing to its high protein content (50%-70% protein of its dry weight) and broad

spectrum of other nutrients viz. minerals, vitamins, lipids, carbohydrates, pigments and other trace

elements [82]. Tavelmout Corp., a biotech company based in Japan developed a closed flat panel

photobioreactor system which enhances protein productivity in Spirulina about 20 times higher than

that of soybeans [83]. While, Siva Kiran et al. [84] reported that the protein content in Spirulina is

higher compared to other foods such as milk, chicken, beef, and some plants. In food application,

Spirulina and Chlorella or its protein were incorporated in various types of food, such as milk-based

products, bread, biscuits, instant noodles and pasta to produce protein-enriched functional food [85-

88].

The extraction of protein from microalgae comprised steps including cell disruption, extraction

and product purification; cell disruption techniques involve mechanical action (high-pressure

homogenisers, bead mills), ultrasounds, enzymatic or chemical treatments, thermal or osmotic shocks

(repeated freezing/thawing) [89]. For efficient protein recovery (76%) from Chlorella sp., Ursu et al.

[90] suggested an alkaline treatment followed by isoelectric precipitation. Meanwhile, Chia et al. [91]

proposed an effective approach using ultrasound-assisted three phase partitioning method for

efficient protein extraction and an optimised conditions for high protein recovery which is applicable

for the future integrated bio-separation technique for biomolecules extraction from microalgae as

well as to improve the current downstream bioprocessing techniques.

Chlorella and Spirulina are known for its high-quality protein attributed by the well-balanced

amino acid composition according to FAO/WHO recommendation [92], digestibility coefficient, as

well as by its biological value of the amino acids absorbed from the food [93]. Nevertheless, interest

to acquire good quality protein for human nutrition continues with exploration on different strains

of microalgae, as such, a study on Australian microalgae species in James Cook University/MBD

Energy Research facility using Scenedesmus sp., Nannochloropsis sp., Dunaliella sp., and a designed

freshwater chlorophytic polyculture (CPC; consisting of Schroederiella apiculata, Scenedesmus

pectinatus, Tetraedrom minimum, Mesotaenium sp. and Desmodesmus sp.) exhibited high quality protein

in all studied microalgae. The protein was suitable for human consumption which was determined

and supported by Essential amino acid index (EAAI) and was comparable to the commercial Spirulina

and Chlorella products. The Australian strains earned higher score for EAAI due to presence of higher

essential amino acids such as histidine, phenylalanine, threonine and lysine. In addition, the selected

strains also displayed a comparable nutrient strength and taste as the Spirulina and Chlorella species

respectively, thus, suggesting the potential of these microalgae for future commercialization in

human nutrition area [94].

54.2 Vitamins and minerals

Microalgae constitute important source of almost all vitamin and essential minerals. Spirulina

was reported as rich source of vitamins B1, B2, B12 and high content of amino acid up to 62%, [95], all

of which have facilitated its claim as superior to other microalgae [96]. Seghiri et al. [95] also suggested

that Spirulina benefits may be attributed by the presence of macro-minerals as well as the trace

elements. Various studies have shown that Spirulina as a good source of pro-Vitamin A, in term of

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 16: Review Isolation of Industrial Important Bioactive

16

beta-carotene and exhibited better effects than the synthetic Vitamin A due to its good bioavailability

[97,98]. This species also was reported for high content of Vitamin B12, in the range of 120-244 µg per

g dry weight, however, studies suggested that most of it were pseudo B12, an analogue which has a

similar structure and bound to specific B12 transporter but does not exert any health benefit [65,99].

In contrast, the latest study by Madhubalaji et al. [100], provided a scientific validation for the use of

Spirulina as Vitamin B12 source. However, more studies needed to substantiate Spirulina as a viable

source of Vitamin B12 in human. In view of high bioavailability of iron from Spirulina, Puyfoulhoux

et al. [101] concluded that Spirulina constitute adequate source of iron for human consumption.

Chlorella is another commercial species with a rich source of proteins, essential amino acids,

vitamins (B‐complex, ascorbic acid), minerals (potassium, sodium, magnesium, iron, and calcium).

Chlorella sp. is suggested to be the best candidate for vegan source of Vitamin B12 and used in food

supplement [102]. According to Kumudha et al. [69], Vitamin B12 detected in the Chlorella vulgaris

present as methylcobalamin, a biologically active form suitable for human consumption while

Merchant et al. [103] reported that participant supplemented with 9 g Chlorella pyrenoidosa daily

mitigated vitamin B12 deficiency in vegetarian and vegan participants. A study by Nakano et al. [104],

suggested 6g Chlorella supplementation significantly reduced the risk of pregnancy-associated

anaemia, proteinuria and oedema. All of these, suggest microalgae enriched food as a good source of

nutritional supplements especially for strict vegetarians owing to its rich high-value nutrients as well

as Vitamin B12. Chronopoulou et al. [105] proposed an improved approach to extract fat-soluble

vitamins by using supercritical CO2 while a previous study compared six extraction methods for

Vitamin B12 from Spirulina and found the aqueous extraction suits best for this studied compound

[106].

4.3 Fatty acids

Microalgae contain distinctive profile of lipids especially the fatty acids, for example, EPA or

DHA with feasible commercial value. The ability to produce and accumulate high amount of PUFAs,

makes microalgae even more valuable as nutraceutical. Omega-3 fatty acids known as a good source

of dietary supplements which highly recognised and recommended for its health benefits specifically

in disease prevention [107] and human nutrition. The algal oil often used in liquid or capsule form

and benefits in particular, vegetarians as well as populations with low seafood diets. Application of

algal oils in food achieved popularity after advancement in the microencapsulation and refining

technology which prevents any off-flavour towards the food when combined together with the oils.

The algal oils are enriched or fortified in myriad of foods such as dairy products, nutritional bars,

bakery products and etc. [108] to further increase product nutritional value.

Microalgae produce total lipid up to 30-70% of dry weight depending on the species; which

plays a critical role as energy stockpile during adverse condition and cell division. Fatty acids are

component of complex lipids and the structural variation attributes to their multitudinous benefits.

Based on the polarity of the molecular head, fatty acids can be categorised into two groups; i) neutral

lipids and ii) polar lipids. Fatty acids production can be increased in the microalgae by manipulating

various environmental factors such as oxygen level, temperature, light exposure, pH as well as

limiting the nutrient supplementation [109]. A number of studies have suggested nitrogen starvation

to increase the neutral lipid and triacylglycerides (TAG) synthesis in microalgae [110,111].

Meanwhile, exposure to low light intensity and therefore low temperature was reported to enhance

PUFAs production [112,113]. Chlorella sp. culture in low CO2 showed to promote high contents alpha-

linolenate fatty acid, in contrast Chlamydomonas reinhardtii mutant cia-3 exhibited higher content of

PUFAs in culture condition with high CO2 concentration [114,115].

Methods of harvesting lipid from microalgae include mechanical pressing, homogenization,

milling and solvent extraction. The solvent extraction is imperative to the polarity or/and solubility

of the lipid of interest. Other techniques such as enzymatic extraction, supercritical extraction,

ultrasonic-assisted extraction, microwaves are used to facilitate lipid extraction by solvent [116].

Nevertheless, only a few cell disruption techniques feasible for commercial application, namely,

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 17: Review Isolation of Industrial Important Bioactive

17

steam explosion [117], enzymatic hydrolysis, bead milling and horn sonication [116]. Purification of

algal oil (PUFAs) is an essential step as the crude form of these oils are inedible due to its impurities,

odour, taste and unattractive for consumption as its turbid appearance. Speed of operation and

process condition are critical factors in the oil purification techniques as the oil is sensitive to

oxidation. Manufacturer such as Martek Bioscience Corporation has described the oil recovery and

purification steps of DHA from algae oil. For instance, protease enzyme was used to break the protein

in the cells of Schizochytrium sp. to release the oil into the culture broth forming an emulsion, thereafter

isoprophyl alcohol was added to separate the oil. While for Crypthecodinium cohnii, hexane solvent

extraction method was applied since the enzyme method is incapable to hydrolyse the algae cellulosic

layer. Following the solvent extraction, the cell walls are removed by centrifugation and the oil was

recovered after solvent evaporation [118].

Several studies and reviews have been published on the different algae strains, namely,

Nannochloropis oculata [76,77], C. cohnii, Schizochytrium sp., Ulkenia sp. which are associated with the

presence of high concentration of PUFAs in its lipid [119] and manifested to have better

bioavailability compared to other fish oils [77]. These strains have been commercially used to

develop high purity marine oils by companies such as Qualitas Health, DSM-NP, Lonza and GCI

Nutrients. DHA oil from C. cohnii (40-50%) are fortified in infant formula milk by company such as

Martek, USA, thus, the cultivation and manufacturing processes follow a strict regulation of FDA

and current Good Manufacturing Practice (cGMP). This fortified formula sold in more than 60

countries [3]. Recently, two more strains with potential commercial values namely, Phaeodactylum

tricornutum and Porphyridium purpureum for EPA and other compounds were suggested [120,121].

4.4. Natural pigments

Microalgae are also a rich source of pigments which are used as bio-colourant or food additives

in various products. Natural pigments are highly coveted by food industries as an alternative to

synthetic sources which has various health implication [122]. One of the key players in this natural

pigment industry is a Chinese astaxanthin supplier, Algae Health Science who operates one of the

biggest facilities in the world, using glass tubes to cultivate Haematococcus pluvialis microalgae for

astaxanthin extraction in Yunnan, China (Figure 4) [123].

Figure 4. Closed tube bioreactor algae production facility in Yunnan Province, China (Courtesy:

Nicholas Cheong, Nova Laboratories Sdn. Bhd.).

Closed system cultivation was favoured by this company as it provides better purity and

prevents contamination, independent of local weather as well as providing consistent production.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 18: Review Isolation of Industrial Important Bioactive

18

Astaxanthin from Haematococcus has been marketed under various name and companies, as such

Cyanotech Corp. from the US commercialized astaxanthin under the name of BioAstin [124],

AstaReal by Fuji Chemical Industry, Japan, Astaxanthin GoldTM by Nutrigold as well as various

patent applications for dietary supplements, health supplements, as antioxidanst, beverage

colourants and other functions [125,126].

Chlorophylls, a photosynthetic green pigment present in microalgae as chlorophyll a (blue-green

colour), b (brilliant green), c (yellow-green), d (brilliant/forest green) and f (emerald green) dependent

on types of algae [127]. Chlorophyll a, which is abundantly found in cyanobacteria such as Chlorella,

has been extensively used as colouring agent due to its stability [128]. Another prominent colourant

from the microalgae with commercial applications is phycobiliproteins, a water-soluble fluorescent

pigment commonly present in cyanobacteria which can be classified as four major groups according

to their colours and light absorption characteristics, namely, phycoerythrin, phycocyanin,

allophycocyanin and phycoerythrocyanin [129,130]. The phycobiliproteins were shown as a strong

antioxidant which contributes to high- value nutraceutical product and its application in food mainly

in dairy products, chewing gums, candy, beverage mixes and ready-to-eat cereals [131,132]. Besides

being a functional ingredient and safe food colourant, this protein (eg. deep blue colour protein C-

phycocyanin) are used as dietary supplement coating and functional food additives [88].

Since the global demand of natural colourant especially astaxanthin is growing at an

unprecedented rate mainly for food supplements, thus, their average prices in the market are in the

range of 2,500 USD per kilograms and its annual worldwide worth of 200 million USD [133]. Species

like Dunaliella ascribed to its high beta-carotene contents, contributes a total turnover of about 75

million USD where 60 million from it contributed by food supplements [67].

Many reviews and research papers offer detailed steps on how to improve and enhance related

algae cultivation and extraction methods. Development of spiral-tube photobioreactor to cultivate D.

salina contributes to an enhanced and continuous cultivation for high output of β-carotene [134].

Additionally, factors such as high light intensity, temperatures, nutrient limitation and high salt

concentrations increase the β-carotene production [135]. A recent study by Xu and Harvey [136],

suggested the use of high-intensity red light with sufficient nutrient content for high carotenoid

production by up-regulating the whole biosynthesis pathway of carotenoids. Besides its applications

as a food colourant, beta-carotene is used as an additive in multivitamin supplement and tablets [3].

Harvesting and extraction of certain algal strain can be challenging due to its nature, as such D.

salina harvesting deemed to be most difficult and costly compared to other commercial strains

attributed to its lacking of cell walls, low concentration and small size. Many approved techniques

have been developed to tackle the problem such as centrifugation and flocculation, a patented

method involving mechanical harvesting [137] while Pirwitz et al. [138] suggested centrifugation

without flocculation method which was identified as most cost-effective technique. Other commonly

used extraction methods for pigments such as astaxanthin are by mechanical treatments, chemical

treatments, pressurized extraction, ultrasound and microwaves among others while a few researches

have taken a greener option by using non or less toxic solvents such as acetone and ethanol for the

extraction process [71]. Molino et al. [70] suggested a mechanical pre-treatment before accelerated

extraction using green solvents.

A continuous search for potential new strains for large- scale production of these pigments are

still advancing as the current major strains are still limited to a few species, namely, Spirulina

platensis, Haematococcus pluvialis and Chlorella sp. Singh et al. [139] discovered Asterarcys quadricellulare

PUMCC 5.1.1 strain, a green microalga with a promising characteristic for carotenoid production

while Kaushal et al. [140] reported a new cyanobacterium Nodularia sphaerocarpa PUPCCC 420.1 which

produces phycobiliprotein and reckoned as a good candidate for production at commercial scale.

5. MICROALGAE IN COSMECEUTICALS

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 19: Review Isolation of Industrial Important Bioactive

19

Skin entity is the largest organ which acts as a physical barrier to protect the human body from

harmful external agents, and is composed of three main layers (epidermis, dermis, hypodermis) [141-

143]. A healthy and radiant skin is maintained via a balance between synthesis and degradation of

the matrix proteins such as collagen, elastin and glycosaminoglycans (hyaluronic acid). The

corresponding proteinases coupled with continual synthesis are the maintenance mechanism

involved in this process. However, this balance is affected via both chronological as well as photo-

ageing by coupling the proteinases up-regulation, where this enhances protein degradation along

with synthesis down-regulation. Protein synthesis simulation or alteration of their breakdown via

proteinases is the suggested mechanism of action for improving this inequality [144].

The synthesis of matrix proteins or proteinases inhibition are attempted with the application of

a wide range of specific skin formulation products containing compounds such as ethanolamines,

sodium lauryl sulphates, polypeptides or oligopeptides. However, the possibility of developing

allergic reactions and other adverse conditions led to continuous search for natural bioactive

compounds that could be applied in cosmetic formulations. As such, bioactive compounds

originating from microalgae have gained much popularity of its amazing moisturizing, thickening,

pigmenting, anti-ageing, skin whitening, sunscreen protection, and many other properties [143,144].

In a review by Mourelle and colleagues, cosmetics are defined as products aimed for improving

skin appearance, structure and morphology in assistance of excipients and active ingredients

specifically suited for various skin types [145]. Several microalgae species extracts are widely applied

in cosmetic-based industries, especially for skincare products [146]. These include face and skincare

(anti-ageing cream, emollient, anti-irritant in peelers, refreshing care, sunscreens) as well as hair care

products [146]. Among the common microalgae species include Arthrospira sp., C. vulgaris, D. salina,

S. platensis, Chondrus crispus, Mastocarpus stellatus, Ascophyllum nodosum, Alaria esculenta and N. oculate

[147]. In cosmetic industry, Cu-Chl (CI 75810) is applied for use in hair colour products, colour

cosmetics and bleaching products and categorised as non-toxic [143]. Figure 5 highlights the main

potential applications of various microalgae in cosmetic industry, accounted by its metabolites or

bioactive compounds, which are described in detail below.

Figure 5. Potential applications of microalgae in cosmeceuticals.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 20: Review Isolation of Industrial Important Bioactive

20

5.1. Anti-ageing

Ageing is a process that involves a decrease in skin structural proteins (elastin, collagen,

hyaluronic acid) synthesis leading to loss in skin elasticity, laxity, integrity and finally gives rise to

visible signs of ageing. Apart from decreased protein synthesis, age-dependent down-regulation of

skin elasticity is also contributed by an up-regulation of the corresponding proteinases. Hence, the

most appropriate anti-ageing strategy in order to attain smooth and healthy skin is by controlling the

skin structural constituents’ degradation via proteinase activity regulation. In addition, frequent and

repetitive exposure to ultraviolet (UV) results in over-production of reactive oxygen species (ROS)

that induces oxidative cellular stress and promotes genetic alterations thereby affecting matrix

protein structure as well as functions, which eventually results in skin damage. In order to overcome

this issue, there is a strong need to scavenge these free radicals [144].

β-carotene, is one of the microalgae pigments commonly synthesized by D. salina that assists in

the prevention of free radicals that initiate premature ageing [143,148,149]. Another carotenoid

compound, lutein is also reported for its protective reactions towards UV radiation and can be

obtained from several microalgae (Scenedesmus salina, Chlorella, C. vulgaris, Scenedesmus obliquss, D.

salina and Mougeotia sp. [143,150]. Lycopene is also a microalgae carotenoid pigment that is employed

in skincare products for its anti-ageing properties where this compound neutralizes oxygen-derived

free radicals [143,145,151].

5.2. Antioxidant/Stress protection

Oxidative stress process impacts the skin mainly in terms of premature ageing, uneven skin

tone/texture and also might disintegrate the essential proteins. Once the collagen and elastin in the

dermal skin layer are diminished via oxidative stress, this also involves significant DNA damage,

inflammatory response, reduced antioxidant protection and the generation of matrix metalloprotein.

In the long run, this results in expedited ageing process along with significant appearance of wrinkles

with loss of elasticity [152]. Fucoxanthin, that is predominantly present in Phaeodactylum tricornotum,

Odontella aurita and Isochrysis aff. Galbana is one important metabolite that was identified to portray

antioxidant activity as well as preventing oxidative stress. Astaxanthin is reported to possess higher

antioxidant activity compared to vitamin A or E, and a recent study by Davinelli and co-workers in

the year 2018 highlighted its potential as a potent anti-wrinkle and antioxidant agent [153].

5.3. Free radical/Sunscreen protection

UV radiation is reported to be beneficial in limited duration; however, prolonged exposure is

not advisable as this will result in severe skin damage. In order to prevent these harmful effects,

various skincare products are usually applied especially by women that could range from sunscreens,

sunblock lotions or anti-ageing serum. Microalgae-derived carotenoid pigments such as astaxanthin,

lutein, zeaxanthin and canthaxanthin found abundantly within Dunaliella and Haemotococcus sp. are

noted for their protective properties against extensive sun damages [154]. In addition, orange-

pigmented violaxanthin compound isolated from Nannochloropsis oceania is proven to significantly

block UVB-detrimental effects along with decreased cell viability and increased ROS production

[155]. Fucoxanthin, is another microalgae-derived pigment that has been shown to impart protective

effect against sunburn [156].

5.4. Pigmenting agent

Almost all cosmetic products are incorporated with synthetic colourants which attracts much

attention of the cosmetic industry towards identifying colourants from natural sources for long-term

sustainability. Microalgae pigments fit well into these criteria where it is classified as a naturally

sustainable source with added health benefits that include UV protection, anti-ageing, antioxidant

and anti-bacterial properties. Carotenoids, chlorophylls and phycobiliproteins encompass the major

classes of pigments in microalgae; where they impart various colours ranging from green, yellow,

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 21: Review Isolation of Industrial Important Bioactive

21

brown and red making it suitable as an alternative to synthetic colourants [157]. Chlorophylls are

extensively applied for use as cosmetic colourants, along with its other use in deodorants as an odour-

masking agent. Chl a, predominantly found in Chlorella and Spirulina sp. is a blue-green compound

and widely used for its stable nature [145]. In contrast to chlorophylls, astaxanthin gives rise to strong

red pigmentation, that are highly useful in most cosmetic products [158]. Phycoerythrin exerts intense

pink fluorescence and is formulated for use in lipsticks, eye shadow, and other make-up essentials.

Phycocyanin, in contrast, imparts blue fluorescence and is sourced from cyanobacteria (Spirulina sp.

or Arthrospira sp.). In the Japanese cosmetic market, phycocyanin isolated from Arthrospira sp. has

been commercialized as cosmetic colourants and eye shadow products [143].

Apart from being used as colourants, these microalgae-pigments are also utilised in tanning

pills. Tanning, is a process defined as the skin adaptation to UV exposure whereby the increased

melanin level plays role in protecting the skin from sunlight rays that induce free radical formation

[143,159]. Cantaxanthin, a pigment that is found predominantly in green microalgae (Chlorella sp.,

Haemotococcus sp. Nannochloropsis sp.) is one of the most common ingredients utilised in tanning pills.

Canthaxanthin imparts its tanning effect by darkening the skin via deposition of its red-orange colour

in the epidermis and subcutaneous fat. However, it is important to address that canthaxanthin

tanning pills are yet to be approved by FDA where adverse effects of urticarial, hepatitis and fatal

aplastic anemia have been noticed with daily consumption of these pills [143].

5.5. Whitening agent

In contrast to tanning effect, the favourability towards fairer skin tone especially among Asian

women has attracted much interest on whitening products as part of their beauty regime. Whitening

effect of skin arises with the inhibition of tyrosinase enzyme, which plays vital role in melanin

biosynthesis. Tyrosinase catalyses melanin synthesis via L-tyrosine hydroxylation to 3,4-dihydroxy-

L-phenylalanine (L-DOPA) as well as oxidation of L-DOPA to dopaquinone followed by further

conversion to melanin [160]. Several microalgal species are noted for their superior tyrosinase

inhibition activity, such as N. oculate or H. pluvalis. Both zeaxanthin and astaxanthin pigments from

these microalgae species are addressed for their anti-tyrosinase ability, making them suitable for

cosmetic products intended for skin whitening [143,161,162].

5.6. Moisturising agent

Moisturisation is a very critical step in preventing premature ageing of the skin, by maintaining

its elasticity and radiance. The level of moisture retained within the human body is correlated with

regulated water transport, corneocytes and hyaluronic acid as well as washing frequency [163.164].

In common practice, cosmetic range of products intended for moisturising effect is formulated with

hydroxy acid however; this elevates the price range due to its limited supply. As such, microalgae-

derived polysaccharides are preferred for its abundance as well as environmental-friendly nature. A

review by Wang and colleagues highlighted the potential of an extract from Chlorella vulgaris that

functions to support skin tissue along with collagen synthesis that assists in reducing wrinkle

formation [164].

5.7. Anti-inflammatory

Another vital and challenging issue of cosmetic products lies on the effect of neurogenic

inflammation that imparts its implications of irritation and itching as the adverse effects. It is

worthwhile to understand that most skin metabolic processes are affected by numerous signals in

response to stress-induced brain nervous stimulus [165]. These signals include secretion of different

hormones/substances from the various skin cells, such as indicated in Table 4 below.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 22: Review Isolation of Industrial Important Bioactive

22

Table 4. Anti-inflammatory secreted signals of the different skin layers.

Skin cells Secreted compounds Reference

Keratinocytes, melanocytes Corticotropin-releasing hormone (CRH), adrenocorticotropic hormone (ACTH), catecholamines

[165-167]

Dermal fibroblasts ACTH, cortisol, prolactin

Skin nerve endings Adrenaline, noradrenaline, substance P

Sebaceous glands CRH, prolactin

Cutaneous nerve endings and almost all skin cells

Produces and responds to special cytokines (neurotrophins) [165,168,169]

In response to nervous stress, the neurotrophins signal becomes crucial for some inflammatory

process (atopic dermatitis, psoriasis) where it induces proliferation of cutaneous nerve endings along

with symptoms ranging from itchiness and pain [165,170,171]. Despite of the limited findings

available on microalgae extracts implications on skin disorders, however, there are several bioactive

compounds derived from microalgae that are shown to be effective. One such example is of the

compound astaxanthin where it is vital in suppressing neuropathic pain in rats; where this involves

inhibition of N-methyl-D-aspartate receptors that is also part of pain mechanism caused by

neurogenic inflammation [165,172,173].

5.8. Industrial applications of microalgae in cosmeceuticals

Microalgae-based products have gained much popularity, and are continuously ventured.

Although the marketed products keep increasing over the years, yet it is still under par compared to

its active ingredient potential. Most of the products are those focused mainly on antioxidant and

photo-ageing protections; where there are still much more to be explored in terms of products

intended for skin appendages treatment, modulation of fat adipokines as well as the skin microbiota

[165]. At the same time, it is still important to highlight that it is economically disadvantageous in

microalgae exploitation as an isolated active compound source due to the relevantly high cost of

biomass production as well as purification [3,165,174,175]. Despite of these, continual

recommendation and exploration to utilise microalgae-derived metabolites for the synthesis of

products with multiple uses are inter-related with it being fully natural source along with its

numerous benefits especially in cosmetic care [165,176].

In a review by Mourelle and colleagues, it is highlighted that several cosmetic-based industries

have invested in establishing their own microalgae cultivation. These include renowned companies

such as Louis Vuitton Moët Hennessy (Paris, France), Danial Jouvance (Carnac, France) and AGI

Dermatics (America). One of the latest highlighted products in 2020 is the Argan Beta-Retinoid Pink

Algae Serum launched by Josie Maran. This product is comprised of the main compound β-carotene

extracted from pink algae that play vital roles to eliminate fine lines, wrinkles, dark spots, dullness,

dry and flaky skin. The serum is also enriched with organic argan oil, quercetin and is claimed to be

free of synthetic colours, fragrance, parabens, petrochemical or phthalates [177]. In addition, various

other microalgae-rich extract products are also commercialized on a continual basis. Among these,

several products available in the market are as shown in Table 5 below.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 23: Review Isolation of Industrial Important Bioactive

23

Table 5. Selected representation of the commercialized microalgae-derived cosmetic care products. 1

Microalgae Commercial name/Company Application Reference

Dunaliella salina Blue RetinolTM Stimulates skin cell growth and proliferation [178]

Nannochloropsis oculata Pentapharm Promotes excellent skin-tightening [145,178]

Chlorella vulgaris Dermochlorella/CODIF Recherche and Nature Stimulates collagen synthesis, supports tissue regeneration, reduces wrinkle

formation, anti-ageing

[3,179]

Arthospira sp. Protulines, Exsymol Assists in skin-tightening and repairs signs of ageing skin [147]

Phaeodactylum tricornutum Givaudan Prevents premature ageing [145]

Argan Beta-Retinoid Pink Algae Serum Josie Maran Eliminate fine lines, wrinkles, dark spots, dullness, dry and flaky skin [177]

Porphyridium sp. Alguard®/Frutarom Anti-wrinkle, protection from UV damage [154,180]

Porphyridium cruentum SILIDINE®/Greentech Improves skin aspect, decrease redness, vascular toning [145]

Cicatrol®/Greensea Anti-ageing, antioxidant [181]

Tetraselmis suecica O+ Gold Microalgae Extract®/Greensea Anti-inflammatory [181]

Scenedesmus rubescens Pepha®-Age/DSM Nutritional Products Protects from photo-ageing, pro-collagen [182]

Nannochloropsis oculata Pepha®-Tight/DSM Nutritional Products (Pentapharm) Protects from oxidative stress, simulates collagen synthesis and skin

tightening

[183]

Dunaliella salina Pepha®-Ctive/DSM Nutritional Products (Pentapharm) Stimulates collagen synthesis, cell proliferation, energy metabolism

Dunaliella salina, Haematococcus pluvialis REVEAL Color Correcting Eye Serum Brightener/Algenist Treats uneven and dull skin, dark circles under eye, gives brighter complexion [184]

Chlorella vulgaris Phytomer/Phytomer Protects skin, neutralises inflammation [154]

Chlorella sp. Golden ChlorellaTM/Terravia Holdings Hydrates skin and hair [145]

2

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 24: Review Isolation of Industrial Important Bioactive

24

Apart from the commercialized products, a recent study in 2017 reported on the potential of a

novel, non-fastidious freshwater microalgae (Chlorella emersonii KJ725233) for possible application in

cosmeceuticals accounted by its anti-ageing, antioxidant and anti-inflammatory properties. Its ability

to inhibit elastase, hyaluronidase, collagenase and to scavenge free radicals supports skin tissue that

gives rise to skin rejuvenation. In addition, phytol compounds, shown to reduce the proteinases

activity and thus, reduces the dermis inflammation [144].

6. MICROALGAE IN BIOFUELS AND ENERGY

Energy production has become the priority of the world’s population due to the fuel demand

for transportation, electric power generation and operation of manufacturing plants. There are two

groups of global energy sources; non-renewable (fossil) and renewable. The renewable energy

sources which have been investigated to displace the depleting fossil energy sources include solar,

wind, nuclear, geothermal, hydrogen, waves, tidal and algal biomass energy [185,186].

The first generation of biofuels comes from terrestrial crops (e.g. maize, sugarcane, sugar beet

and rapeseed) which has many drawbacks/limitations causing destruction of forests and the

abundant usage of water. The second generation is from forest residues, lignocellulosic agriculture

and from non-food crop feedstocks which disadvantage is the land use [187,188]. Biofuels are the

renewable energy sources produced from algal biomass which is the third generation of biofuels and

is a promising alternative method to the previous methods. The advantages of using algae to generate

fuels/energy are that it can be produced all-year-round, easy to cultivate in water without much

effort, requires less water compared to terrestrial crops, does not require pesticides or herbicides thus

reducing the cultivation cost [189,190].

Alga biofuel has the potential to meet global demand since they do not compete with the

production of food products. They can be exploited to biofuels inclusive of biodiesel,

bioethanol/biobutanol, syngas, biochar, biogas and energy (electricity/heat) through various

processing technologies such as thermochemical conversion (transesterification, thermochemical

liquefaction, direct combustion, gasification and pyrolysis) and biochemical conversion

(fermentation, anaerobic digestion and photobiological hydrogen production) (Figure 6) [16,191].

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 25: Review Isolation of Industrial Important Bioactive

25

1

2 3

Figure 6: Various methods of producing biofuels/energy from microalgae biomass. Adapted from [16,191]4

Microalgae Biomass

BiochemicalConversion

Transesterifi-cation

Biodiesel

Fermentation

Bioethanol/

Bobutanol

Anaerobic Digestion

Biogas/

Methane

Biophotolysis

Hydrogen

Thermochemical Conversion

Gasification

Syngas

Thermochemical

Liquefaction

Bio-Oil

Pyrolysis

Bio-Oil

Syngas

Charcoal

Direct Combustion

Energy

Electricity

Heat

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 26: Review Isolation of Industrial Important Bioactive

26

6.1. Biochemical Conversion

Biochemical conversion is the conversion of microalgae biomass to biofuels with the aid of

biocatalysts, such as enzymes and microorganisms such as bacteria and yeast [192]. This method

works well with high-water content biomass. The biochemical conversion includes

transesterification, fermentation and anaerobic digestion.

i)6.1.1 Transesterification- Biodiesel

Fatty acid methyl ester (FAME) called as biodiesel is produced by transesterification or

esterification; a chemical reaction between alcohol and triacylglycerides without or with catalysts

such as alkaline catalyst (sodium hydroxide or potassium hydroxide) or acid catalyst (sulphuric,

sulphonic, phosphoric, and hydrochloric acids) [193,194]. Many species of microalgae such as

Scenedesmus obliquus, Neochloris oleabundans, Nannochloropsis sp. and C. vulgaris contain a high content

of lipids including triacylglycerides that are needed for the biodiesel production [195,196]. The

production of biodiesel involves several stages including lipid extraction using solvents; methyl or

ethyl alcohol (unrefined lipid), followed by lipid recovery with centrifugation-solvent evaporation

(purified lipid and free fatty acids), transesterification or esterification (biodiesel production). The

final step of this process is the removal of the solvent for the biodiesel recovery [196]. Biodiesel from

renewable resources such as microalgae has several advantages compared to petroleum diesel; non-

toxic, biodegradable, renewable and contains low levels of carbon monoxide, soot, hydrocarbons and

particulates. It also discharges a low level of CO2 of up to 78% lower [197]. Many researchers have

look into the possibility of producing biodiesel alongside with bioethanol from algae biomass. A

study by Wang et al. [198] had shown encouraging results of Tribonema sp. with enriched lipid and

carbohydrate content for biodiesel and bioethanol production. The conversion rate of lipid extracted

with hexane-ethanol to biodiesel was 98.4% and the maximum yield of ethanol using yeast

Saccharomyces cerevisiae was 56.1%.

ii)6.1.2 Fermentation- Bioethanol

The chemical and enzymatic pre-treatment method are usually used in the production of

bioethanol. The common strong acids used include hydrochloric, sulfuric and nitric acid, whereas

enzymes such as amylases, cellulases and invertases are used. However, using enzymes can be rather

expensive if involves large scale production [199]. Sometimes the cell walls are disrupted using a

mechanical method such as by 24/24 freezing/defrosting cycle to release the intracellular substances

[200]. The fermentation of the treated biomass is usually using yeast Saccharomyces cerevisiae or

bacteria at warm temperature (30-40 oC) to obtain ethanol and a bioreactor is later used for up-scaling

[201,202]. The quality and quantity of the produced bioethanol are fully determined by the method

used and fermentation process parameters inclusive of pH, temperature and the fermenter organism

used [203].

Previous studies have shown that several microalgae species can act as an effective feedstock for

the production of bioethanol. The species include Chlorococcum humicola, Scenedesmus abundans PKU

AC 12, Chlorella vulgaris FSP-E, Scenedesmus obliquus CNW-N, Porphyridium cruentum, Desmodesmus

sp, Spirulina platensis and Scenedesmus acuminatus (Refer Table 6). A comparison study of red

microalgae (P. cruentum) culture conditions for bioethanol production was done by Kim and co-

workers in 2017 [204]. The results indicated ethanol conversion yields from P. cruentum cultured in

freshwater was much higher (70.3%) than the P. cruentum cultured in seawater (65.4%). Recently,

Chandra et al. [205] study the effect of the cultural variables on the carbohydrate accumulation of

Scenedesmus acuminatus to produce bioethanol. The bioethanol yield obtained was 0.12 g/g with the

supplementation of lysine in the medium culture and at higher initial culture pH (pH 9.0).

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 27: Review Isolation of Industrial Important Bioactive

27

Table 6. Bioethanol production from various microalgae species.

Microalgae Ethanol yield (g ethanol/g substrate) Reference

Chlorococcum infusionum 0.26 [206]

Chlamydomonas reinhardtii 0.24 [207]

Chlorococcum humicola 0.52 [208]

Scenedesmus abundans PKU AC 12 0.103 [209]

Chlorella vulgaris FSP-E 11.66 (87.59%) [210]

Scenedesmus obliquus CNW-N 8.55 (99.8%) [211]

Desmodesmus sp. FG

strain SP2-3 (unidentified green microalga)

0.24 [212]

Scenedesmus acuminatus 0.12 [205]

iii)6.1.3 Anaerobic Digestion- Biogas/methane

Microalgae can be converted to biogas through anaerobic digestion which is a biochemical

process that mineralizes organic material through the action of microorganisms in the absence of

oxygen [213]. Biogas is mainly made up of a mixture of methane, CH4 (50-70%), CO2 (30-45%) and

traces of other gases such as hydrogen, H2 (<2%) and hydrogen sulphide, H2S (<3.5%). The process

involves three stages; hydrolysis, acetogenesis/acidogenesis and methanogenesis [214]. Biogas

production is often reduced due to the thick and rigid microalgae cell wall, therefore it is normally

incorporated with production of other bioproducts such as biodiesel or bioethanol [215]. To

overcome the cell wall problems, pre-treatment is usually done to enhance the process. Pre-treatment

with heat was found to enhance the production of biomethane from Chlorella sp. by 11% as the

temperature was increased to 90°C from 70°C for 0.5h compared to the control [216].

iv)6.1.4 Biophotolysis- Hydrogen

Green microalgae and cyanobacteria can produce hydrogen through biophotolysis process.

They possess chlorophyll a and the photosynthetic systems; Photosystem II (PS II) and Photosystem

I (PS I), that enables them to perform photosynthesis by absorbing solar energy and converting water

into hydrogen and oxygen [217,218]. Biophotolysis can be divided into two; direct and indirect

pathways. During direct biophotolysis, water splitting at PS II generated electron and proton, giving

rise to H2 in both green algae and cyanobacteria. For indirect biophotolysis, the degradation of carbon

compounds will generate protons and electrons for the production of hydrogen which are mostly

found in cyanobacteria [219].

There are several green microalgae and cyanobacteria such as Tetraspora sp. CU2551, R. rubrum,

R. spheroides, Rhodobacter capsulatus, C. vulgaris, C. reinhardtii, Anabaena sp. and Nostoc sp. which have

been widely studied for the hydrogen production [220-223]. The model microalga that is usually used

for research is the C. reinhardtii since some of its organelles such as mitochondrial, nuclear genomes

and chloroplast have been successfully sequenced [224].

6.2. Thermochemical Conversion

Thermochemical conversion is a process whereby microalgae biomass is converted to biofuels

with the involvement of heat at a different temperature depending on the final product desired. This

method can be used for both dry and wet biomass. Normally, in the thermochemical conversion no

chemicals are added and the period for producing biofuels from this method is shorter than the

biochemical conversion [225]. The thermochemical conversion methods include gasification,

thermochemical liquefaction, pyrolysis and direct combustion.

6.2.1 Gasification- Syngas

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 28: Review Isolation of Industrial Important Bioactive

28

Syngas or synthesis gas is mainly composed of methane, hydrogen, carbon monoxide and

carbon dioxide that can be used to generate heat or electricity [185,226]. Microalgae biomass is

converted to Syngas through a gasification process which utilises partial oxidation with a mixture of

oxygen and steam at high temperatures (>700°C) [227].

6.2.2 Thermochemical Liquefaction- Bio-Oil

Crude bio-oil is produced by thermochemical liquefaction in the presence of a catalyst at a

temperature between 300-350°C [228]. Hydrothermal liquefaction is the preferred method for wet

algae biomass which uses water and elevated pressure [227]. The thermochemical liquefaction yield

of microalgae depends on various parameters including type of catalysts and reaction temperature

[229].

6.2.3 Pyrolysis- Bio-Oil, Syngas, Charcoal

Bio-oil, syngas and charcoal can be produced through pyrolysis; the conversion of microalgae

biomass in the absence of oxygen at medium to high temperatures (400–600°C) [230]. Different

products produced at different stages of temperature and duration of the process. Bio-oil is produced

with a moderate temperature (500oC) and a short-time exposure to vapour known as flash pyrolysis.

Syngas is produced through fast pyrolysis at moderate temperature (500oC) and moderate time

exposure to vapour. Charcoal is produced through slow pyrolysis at a lower temperature (400oC) and

long-time exposure to vapour [231].

6.2.4 Direct Combustion- Energy (Heat/Electricity)

Direct combustion is the burning of the microalgae biomass to generate heat, energy or

electricity. It is normally done in a boiler or steam turbine at a temperature above 800oC in the

presence of oxygen. The process can produce 300 MW for domestic or a large-scale industrial process

[16]. However, the overall operation cost may increase due to the pre-treatment requires by the

microalgae biomass such as drying [232].

6.3. Industrial applications of microalgae in biofuels

The potential of microalgae to be used as a source of biofuel in the future to replace the fossil

fuel has always been a concerned [233]. However, there are several challenges that need to be

addressed to enable success which include the development of low-cost, effective cultivation systems,

efficient and energy-saving harvesting techniques because there are problems to discard large

volumes of water during harvesting and methods for oil extraction and conversion that are

environmentally friendly and cost-effective by selecting the best strain for higher production of

biofuel [186,234]. In terms of cost, the harvesting process may involve a lot of money and

commercialization will be too expensive as reported by the U.S. Department of Energy (DOE) who

has been doing research in this area for about 16 years from the year 1980s to 1990s [235]. The only

oil major company known to be still investing in the microalgae biofuel project is the ExxonMobil. In

2017, after 8 years of collaboration with a biotechnology company Synthetic Genomics, they

announced a major breakthrough in producing microalgae with 40 percent more lipids while

maintaining growth rates using CRISPRCas9 genome editing technique [236]. However, to date no

recent developments are known regarding this issue. Although the microalgae-based biofuel

production could not be commercialised in near future, there is still a possibility of using microalgae

for biofuels in the future because of the abundance of resources such as the significant amount of

land, water and CO2 available to support the algal biofuel technology [235]. Lately, algal biofuel

production has attracted interest again from many researchers to explore in this field to be

commercialised later. The areas of study and development that need to be considered to achieve

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 29: Review Isolation of Industrial Important Bioactive

29

success are in the cultivation, harvesting and extraction methods where the growth rate of microalgae

need to be enhanced with a more robust alga growing systems, new species/strains with high lipid

content need to be sought or the lipid content in the microalgae needs to be enhanced in various ways

which are found to be appropriate and the development of an efficient method for the biofuel

extraction needs to be explored so as not to increase the production cost [235].

7. MICROALGAE IN BIOFERTILISERS

Biofertilisers are natural substances or products containing live microorganisms that enhance

the chemical and biological properties of the soils, revive soil fertility and stimulate plants growth

[237]. Plants need nitrogen to grow and the lack of such component can be overcome by giving

fertiliser at an adequate rate. However, excessive and prolonged use of chemical or synthetic

fertilisers resulted in environmental pollution that will eventually cause an imbalance in the

ecosystem [238]. As an alternative, microalgae have been extensively studied to see its potential as

plant biofertilisers and also biostimulants.

The majority of cyanobacteria can fix nitrogen from the atmosphere and several species are

known to be efficient as cyanobacterial-based biofertilisers such as Anabaena sp, Nostoc sp. and

Oscillatoria angustissima [239-241]. Some of the various green microalgae and cyanobacteria species

successfully used as biofertilisers to enhance crops growth include Acutodesmus dimorphus, S.

platensis, C. vulgaris, Scenedesmus dimorphus, Anabaena azolla and Nostoc sp. (Table 7), with Chlorella

vulgaris as one of the most commonly used microalgae in biofertiliser studies. The germination of

Hibiscus esculentus was accelerated using combined seed and soil treated with C. vulgaris. Significant

improvement of the soil nutrient content and microorganism count before treatment with the

biofertiliser was also observed [242].

Table 7. Microalgae species and their use as plant biofertiliser and biostimulant.

Microalgae Crops Reference

Acutodesmus dimorphus Tomato [243]

Spirulina platensis,

Chlorella vulgaris

Maize [244]

Chlorella vulgaris Hibiscus esculentus [242]

Chlorella vulgaris (UTEX 2714),

Scenedesmus dimorphus (UTEX 1237)

Rice [245]

Spirulina platensis,

Chlorella vulgaris

Onion [246]

Chlorella vulgaris Tomato and Cucumber seeds

[247]

Scenedesmus sp. Rice [248]

Anabaena azolla Rice [249]

Nostoc muscorum

Nostoc rivulare

Maize [250]

The treatment using green microalgae/cyanobacteria have shown many beneficial effects on the

plants and soils. The seed germination, plant growth, yield and the nutritional value of the crops is

enhanced besides the improvement of the soil fertility. The carbon and organic content of the soil

were accelerated due to the excretion of carbon (exopolysaccharides) by the green

microalgae/cyanobacteria into the soil and the degradation of the biomass and grazing activity add

on to the increment [251-254]. Studies had shown that those factors influence the microbial activity

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 30: Review Isolation of Industrial Important Bioactive

30

and biomass of other microflora and fauna in the soil which eventually will stimulate the growth of

the crops [251,255].

A study by Bumandalai et al., [247] exhibited the potential of Chlorella vulgaris as biofertiliser for

the germination of tomato and cucumber seeds. The length of the tomato and cucumber roots and

shoots were improved using algal suspensions of 0.17 and 0.25 g/L, respectively. Treatment of plants

with A. dimorphus biofertiliser before seedling transplant showed enhanced germination, increased

production of branches and flowers compared to the control group and the treatment group applied

during transplant [243]. Nayak et al., [248] successfully used de-oiled microalgal biomass of

Scenedesmus sp. as biofertiliser to improve the growth of rice plant.

The cyanobacteria can colonize different parts of a plant tissue such as in the roots and shoots,

stimulating the nitrogen fixation and phosphorus solubilising microbial population on that parts,

thus enhancing and improving the growth, nutritional status and defence mechanism of the plant

and soil fertility [256-261]. Apart from that, cyanobacteria also produce siderophores (organic

compounds) to facilitate in chelating micronutrients (e.g Fe and Cu), known as biomineralization to

make them easily accessible for plants growth [262,263].

Many green microalgae/cyanobacteria are also reported to excrete intracellular hormones to

their surroundings which help to stimulate plants growth [264-266]. Examples are the green

microalgae, Chlorophyta and Cyanophyta [267] and some cyanobacteria strains which produce

cytokinins and auxins which help to stimulate plants growth parameters such as shoot length, root

length, spike length and weight of seeds [268].

The use of green microalgae/cyanobacteria as biofertiliser can increase the activity of defence

mechanism of the plants and improve their immunity by increasing the plant RNA activity,

producing nutrient assimilating enzymes; dehydrogenase, nitrate reductase, acid or alkaline

phosphatase, generating antioxidant and plants defence enzymes such as peroxidase, polyphenol

oxidase, phenylalanine ammonia lyase and β-1, 3-endoglucanase in root and shoot of the plants

[259,269,270]. Different strains of microalgae might provide different levels of defence mechanisms

as shown by a study by Babu et al. [271]. Three different cyanobacteria (Anabaena laxa RPAN8,

Calothrix sp. and Anabaena sp. CW2) inoculated in wheat plant showed the highest activity of

peroxidase, polyphenol oxidase and phenylalanine ammonia-lyase with Calothrix sp. The results

could suggest that using a combination of different types of microalgae with different abilities as

biofertiliser, may contribute to the increase of plants immunity.

Some researchers had tried using a mixture of different microalgae species or a combination of

microalgae and other organic or chemical fertilisers to further enhance its effectiveness. Maize plant

treated with C. vulgaris and S. platensis along with cow dung manure for 75 days under greenhouse

conditions, appeared to improve the growth and yield of the maize plant [244]. Jochum et al., [245]

utilised a mixed culture of C. vulgaris (UTEX 2714) and Scenedesmus dimorphus (UTEX 1237) biomass

as biofertiliser for rice plant. The biofertiliser showed efficacy with a significant increase in height of

the rice plant. Growth parameters of onion plants were found to enhance with the application of a

mixture of S. platensis and C. vulgaris showing higher growth rate and yield compared to the control

group [246]. Consortia of green microalgae and cyanobacteria also showed promising results with

the improved activity of soil microbial, increase in soil organic carbon, macro- and micronutrients

and enhanced plant growth and yield [272,273]. Figure 8 summarises the function of microalgae

biomass as biofertilisers.

7.1. Industrial applications of microalgae biofertiliser

The cost of using chemical fertilisers are increasing every year due to the high demand for foods

from plants since it is one of the major important components for plant growth, especially for food

production. Microalgae-derived biofertiliser can be an alternative for chemical fertiliser and has

gained more popularity due to the natural effects they provide to the crops and the surrounding soil.

The chemical fertiliser may cause several problems, among them are the pollution (soil, water, air) to

the environment, reducing input efficiency, decreasing food quality, developing resistance in weeds,

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 31: Review Isolation of Industrial Important Bioactive

31

diseases and insects, micronutrient deficiency in the soil, degradation of soil and causing toxicity

effect to different beneficial organisms living in the soil [274]. On the other hand, the biofertilisers are

eco-friendly, more economical and more efficient [238]. Several things that are considered before the

commercialisation of microalgae as biofertiliser are the formulation of the inoculate, the nature

application of the products, packaging issues and also products storage [275]. Several countries have

successfully commercialised the microalgae-derived biofertiliser products, which include the

developing and developed countries. Previously in 2019, a Spanish company, Biorizon Biotech, S.L.

had applied a patent for the method of obtaining concentrates of biofertilisers and biostimulants for

agricultural use from green microalgae and cyanobacteria biomass [276]. Among the available

microalgae-derived biofertiliser products in the market are as shown in Table 8.

Table 8. Commercialised microalgae-derived biofertiliser

Microalgae Commercial name Company Country Reference

Cyanobacteria

(Blue-green algae)

Skipper Khad® and

Power Play-90

WSG

Ecological Products

Industries

India [277]

Spirulina sp. Shwe Awzar®

June Industry Limited

Myanmar [278]

Spirulina sp. Algafert® Biorizon Biotech

Spain [279]

Chlorella vulgaris Terradoc® Mikroalg Food and

Agriculture

Industries

Turkey [280]

Cyanobacteria TerraSyncTM

Accelergy Corporation USA [281]

Chlorella sp. EMEK MCT Tarim Ltd Turkey [282]

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 32: Review Isolation of Industrial Important Bioactive

32

1

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 33: Review Isolation of Industrial Important Bioactive

33

8. MICROALGAE IN WASTEWATER TREATMENT

Microalgae can utilise nitrogen, phosphorus, carbon and chemicals such as heavy metals from

waste water as their source of nutrients and this is called as phycoremediation [283]. Conventional

wastewater treatment plants (WWTPs) is designed basically to remove organic matter and nutrients.

Nitrogen (N) and Phosphorus (P) are the most important pollutants in the aquatic environments

because they are discharged into water bodies in huge quantities. Removal of N and P using

microalgal remediation technology has high removal efficiencies which N and P contaminants can

be completely removed from WWTPs, lower costs operation, zero sludge generation and high-value

products can be generated such as fatty acids, pigment, biomass, biodiesel and unicellular protein

[284-287].

New technology in removing heavy metals such as zinc, copper, lead, mercury, chromium and

cadmium from the wastewater (WW) is by using microalgae biomass. Microalgae biomass is an

option owing to low obtaining costs and close to 100% efficiency in the removal of heavy metals

[288,289]. Microalgae also able to remove colorant such as aniline used by textile, cosmetics and food

industries. A major portion of colorant removal is due to the interaction of the molecule with the

reactive sites of the cell wall which permits the elimination of the colorant in a rapid manner [290,291].

Emerging contaminants (ECs) are molecules produced in industries such as pharmaceutical

(used for human beings and livestock) and chemicals, including compounds such as hormones,

antibiotics, plasticizers, antipyretics, antifungal and surfactants which cannot efficiently remove by

WWTPs [292,293]. The removal of ECs is normally done by physicochemical processes such as photo-

oxidation, ozonisation, or oxidation via metallic catalysis. Algae-based bioreactors have gained

special research interest as a promising way to remove pharmaceuticals-based ECs from the

wastewater either partially or completely. A wide spectrum of technologies has been used for the

removal processes for ECs, mainly grouped as physical adsorption, chemical advanced oxidation and

biological degradation processes and combination. Laboratory scale (open pond and bubble column

photobioreactor) showed high removal percentages have been reached for ciprofloxacin, metoprolol,

triclosan, and salicylic acid (>90%); moderate removal for tramadol and carbamazepine (50-90%), and

trimethoprim and ciprofloxacin show fewer promising results with very low removal (<10%) [294].

Based on the latest research, phenol and its derivatives could be metabolized and degraded by

algae. The ability of algae to remove phenolic compounds dependant with the type of algae or

substrate, degradation time and initial pH. However, algae have a certain tolerance range to phenol

[295,296]. Antibiotics are other pollutants that have been released to the environments and create

antibiotics resistance. Recently, microalgae-based technology has been explored as a great potential

alternative for the treatment of wastewater containing antibiotics by adsorption, accumulation,

biodegradation, photodegradation, and hydrolysis [297]. Two freshwater algae species Scenedesmus

obliquus and C. pyrenoidosa were shown to have the capability to transform steroids such as

progesterone and norgestrel [298].

A device called Microalgae-microbial fuel cell (m-MFC) integrated process has been developed

to overcome the problem of fossil-fuel depletion and environmental pollution by generating electrical

energy from wastewater and sunlight, wastewater treatment, CO2 sequestration and biomass

production in single, self-sustainable technology [299,300]. Table 9 showed microalgae used in the

remediation of the pollutants present in the wastewater.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 34: Review Isolation of Industrial Important Bioactive

34

Table 9. Microalgae used in the remediation of the pollutants present in the wastewater.

Pollutants Microalgae bioremediation References

Nitrogen &

Phosporus

Phormidium sp, Spirulina maxima, Chlorella vulgaris, Scenedesmus dimorphus, Scenedesmus quadricauda, Chlorella sorokiniana, Chlorella vulgaris ESP-6 [284-287]

Heavy Metals Chlorella sp., Tetradesmus sp., Scenedesmus sp., Porphyridium sp., Chaetoceros sp., Oscillatoria sp., Sprogyra sp., Scenedesmus sp., Anacytis sp., Chlamydomonas sp.,

Anabaena sp., Ceratium sp., Scenedesmus sp., Calothrix sp., Arthrospira platensis

[288,301,302]

Colorants

Chlorella vulgaris, Cosmarium sp., Nostoc linckia, Spirogyra sp. [303-306]

Emerging Pollutants

Antibiotics

Phenol

Nannochloris sp., Mixture of algae-bacteria consortia in pilot high rate algae pond (HRAP), Mixture of algae-bacteria consortia (dominated by Coelastrum sp.) in HRAP,

Mixture of algae-bacteria consortia in 1-L HRAP S. obliquus, C. vulgaris, Chlorella sp., Scenedesmus sp., Chlamydomonas mexicana, Chlorella sorokiniana

Chlorella vulgaris

Chlorella pyrenoidosa

[307-311]

[296,312]

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 35: Review Isolation of Industrial Important Bioactive

35

9. MICROALGAE IN FEED

Microalgae biomass have also been successfully employed in feed formulations for different

animals like cattle, fish, goat, lamb, poultry, pigs and rabbits. In aquaculture, species of genus such

as Nannochloropsis, Isochrysis, Pavlova, Phaeodactylum, Chaetoceros, Skelotenma, Thalassiosira and

Tetraselmis are used as biomass [313]. A microalgae, D. salina has been applied in animal feed for its

rich protein (57% d.w.) and carbohydrates (32% d.w.) contents [314]. Microalgal based feed affects

the animal's physiology as well as it improves its immune response and fertility. The combination of

5–15% of algal biomass, mixed with animal feed, can be used safely as a partial replacement for

conventional proteins. However, the use of higher concentrations of microalgae biomass results in

less feed intake by some animals due to the lower palatability of the feed. The quality of proteins with

microalgal sources have a comparable or superior functionality (either as foam, emulsifier,

solubilizer, surfactant, or gelling agent) equal or even greater than other commercial protein sources

[315]. Chlorella biomass showed that it can be digested easily up to 5% in the form of paste [44].

Microalgae can provide an important source of essential fatty acid to fish. DHA and EPA are not

synthesized by fish and the only resource to enrich them in fish are by accumulating them in proper

algal paste and store them as fish feed [316,317]. Microalgae, like C. vulgaris and A. platensis, consist

of amino acid profiles that show similarities to soybean which is currently the main source of protein

used in feed. In addition, the microalgal biomass is a potential source of minerals; C. vulgaris, A.

platensis, Micractinium reisseri, Nannochloris bacillaris, and Tetracystis sp. showed very high contents of

iron in contrast to soybean [318].

10. MICROALGAE IN PROTEOMICS

Genetic engineering has been used to manipulate microalgae which allow researchers deducing

metabolic pathways in order to construct new algal metabolism methods. This will enable to produce

new molecules useful for biotechnological applications [319]. Microalgae proteomics can give the

information of how genes of interest are being expressed as proteins which will help researchers to

find proteins with biotechnological applications such as anti-inflammatory proteins [320].

The potential of microalgae in the production of proteins with biotechnological application has

been explored. C. reinhardtii has been extensively studied for therapeutic protein production and

more than 100 such proteins have been expressed successfully in this system [321]. Microalgae-based

biomanufacturing is preferred due to effectiveness in terms of cost and energy, fewer contamination

chances, and simplified downstream processing. Malaria detection protein which are used in ELISA

testing and cell-traversal protein as an antigen from sporozoites and ookinetes of mosquitoes have

been successfully expressed and produced in C. reinhardtii chloroplast [322]. Phycobiliproteins from

cyanobacteria and red algae have been described to present antioxidant, hepato-protective, anti-

inflammatory, immune-modulating, and anticancer effects [51,323-326]. Chlorella pyrenoidosa and

Chlorella vulgaris have shown anti-hypertensive and anti-tumour activities from its isolated peptides

[325].

A protein isolated from Chlorella sorokiniana has high-value bioactive peptides with nutraceutical

and pharmaceutical application, using proteomics techniques [332]. Most of the proteomics

approaches in the area of microalgae research are focused in development of efficient biofuels

production [327-330].

11. CONCLUSION

Microalgae are promising source of generous amount of metabolites with essential health

benefits. Continual exploration for novel microalgae species along with isolation of its bioactive

compounds are in great demand for diverse applications in various fields and industries as raw

material, biomass or high-quality extracts. Despite of the excellent and superior benefits, there are

several limitations that should be addressed as well. The significantly high incurring cost of the

isolation might result in over-priced products, where it is believed this could be partly subsidised

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 36: Review Isolation of Industrial Important Bioactive

36

with utilisation of single solvent to generate high extract yield. In addition, it is essentially important

to highlight that these are natural extracts with unique composition and varies subject to the season,

cultivation condition, as well as the extraction method. Studies are also focusing on extraction of

more than one product from a single biomass for a cost-efficient industry. On the other hand, the

process of bringing a drug candidate to the market requires extensive pre-clinical testing and clinical

trials to determine the safety and efficacy of the drug before it is approved by the FDA and is a very

costly and time-consuming process. Hence, not all potential bioactive will eventually reach the

market and consumers. Indeed, low yield of bioactive and strenuous and expensive purification steps

have been regarded as the main challenge in producing economically viable drugs from microalgae.

Therefore, to overcome these issues some of these active compounds are currently expressed in

suitable vectors to ease purification and hence reduce the purification and production costs. Efforts

on cost reduction by studying effect of different culturing conditions such as light intensity, nutrient

availability and harvesting time on the metabolite yield and bioactivity as well as producing high-

yield microalgae species via genetic modification approaches are currently pursued to accelerate the

commercialization process of some of these bioactives. Meanwhile, advancements in biotechnology

applications can help overcome some issues of poor oral bioavailability or instability in the

gastrointestinal tract of some microalgae bioactives like peptides and fatty acids. This includes novel

approaches using encapsulation technique and nano formulation to improve the solubility and

bioavailability of some microalgae derived compounds for treatment of diseases. Furthermore,

genetic engineering approaches expressing microalgae metabolites in natural organism for drug

delivery have been explored. Nevertheless, more studies especially in in-vivo models and clinical

trials are still needed to determine the safety and efficacy of these novel drugs prior to drug approval

and commercialization.

In conclusion, it is vital that more efforts should be taken in developing multi-functional range

of products, that are affordable and is inter-related with nutritional science that could cater for even

more health benefits comparatively to the ones in current market. These findings would bring much

more insights into the vast potentials of microalgae-derived metabolites that remain to be explored

and assessed. Further roadmap towards enhancing phycoeconomy is recommended by uplifting the

current biological and technology process by taking into account on sustainability and environmental

benefits.

Acknowledgments: The authors thank the Director-General of Health Malaysia and the Director of Institute for

Medical Research (IMR), National Institutes of Health, Ministry of Health Malaysia, for giving the permission to

publish this article. We also thank the staff of Nutrition, Metabolism and Cardiovascular Centre, Institute for

Medical Research, NIH for their continuous support.

Author contributions: Conceptualization: All authors.; Literature review: All authors.; Tables: All authors.,

Writing and review All authors.; Editing: All authors; Revisions and Final editing: All authors. All authors have

read and agreed to the published version of the manuscript.

Conflicts of Interest: The authors declare no conflict of interest.

List of Abbreviations

ACTH Adrenocorticotropic hormone

ALA α-linoleic acid

cGMP current Good Manufacturing Practice

CPC Chlorophytic polyculture

CRH Corticotropin-releasing hormone

CV-N Cyanovirin

DHA Docosahexaenoic acid

DOE U.S. Department of Energy

EAAI Essential amino acid index

ECs Emerging contaminants

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 37: Review Isolation of Industrial Important Bioactive

37

EPA Eicosapentaenoic acid

EAAI Essential amino acid index

FDA Food and Drug Administration

FP-PBR Flat-plate open pond and closed photobioreactor

GRAS Generally Regarded as Safe

HRAP High rate algae pond

IC Inhibitory concentration

L-DOPA 3,4-dihydroxy-L-phenylalanine

m-MFC Microalgae-microbial fuel cell

PBR Open pond and closed photobioreactor

PUFAs Polyunsaturated fatty acids

ROS Reactive oxygen species

SVN Scytovirin

TPBR Tubular open pond and closed photobioreactor

TAG Triacylglycerides

TG Triglyceride

UV Ultraviolet

VCPBR Vertical column open pond and closed photobioreactor

WW Waste water

WWTPs Waste water treatment plants

References

1. Pereira, A.G.; Jimenez-Lopez, C.; Fraga, M.; Lourenço-Lopes, C.; García-Oliveira, P.; Lorenzo,

J.M.; Perez-Lamela, C.; Prieto, M.A.; Simal-Gandara, J. Extraction, Properties, and

Applications of Bioactive Compounds Obtained from Microalgae. Curr Pharm Design 2020,

26, 1929-1950.

2. Norton, T.A.; Melkonian, M.; Andersen, R.A. Algal biodiversity. Phycologia 1996, 35, 308-326.

3. Spolaore, P.; Joannis-Cassan, C.; Duran, E.; Isambert, A. Commercial applications of

microalgae. J Biosci Bioeng 2006, 101, 87-96.

4. Meticulous Research. Algae Products Market by Value (Medium Value, High Value, And

Low Value), Products (Hydrocolloids, Carotenoids, Omega-3 PUFA, Spirulina, Chlorella),

Application (Food and Feed, Nutraceutical, Cosmetics, Chemicals) - Forecast 2022. Available

online: https://www.meticulousresearch.com/product/algae-products-market-forecast-2022

(accessed on 6th October 2020).

5. Doshi, A.; Pascoe, S.; Coglan, L.; Rainey, T.J. Economic and policy issues in the production of

algae-based biofuels: A review. Renew Sust Energy Rev 2016, 64, 329-337.

6. Benedetti, M.; Vecchi, V.; Barera, S.; Dall’Osto, L. Biomass from microalgae: the potential of

domestication towards sustainable biofactories. Microb Cell Fact 2018, 17, 1-18.

7. Sun, Z.; Liu, J.; Zhou, Z.-G. Algae for biofuels: An emerging feedstock. In Handbook of biofuels

production, Elsevier: 2016; pp. 673-698.

8. Enzing, C.; Ploeg, M.; Barbosa, M.; Sijtsma, L. Microalgae-based products for the food and

feed sector: an outlook for Europe. In EUR - Scientific and Technical Research Reports, M, V., C,

P., ER, C., Eds. Publications Office of the European Union: Lxembourg, 2014; pp. 9-18.

9. Mendes, L.B.B.; Vermelho, A.B. Allelopathy as a potential strategy to improve microalgae

cultivation. Biotechnol Biofuels 2013, 6, 1-14.

10. Dragone, G.; Fern, B.; Vicente, A.A.; Teixeira, J.A. Third generation biofuels from microalgae.

In Current research, technology and Educaion topics in applied microbiology and microbial

biotechnology, AM, V., Ed. Formatex Research Center: Badajoz, Spain, 2010; pp. 1355-1366.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 38: Review Isolation of Industrial Important Bioactive

38

11. Tan, J.S.; Lee, S.Y.; Chew, K.W.; Lam, M.K.; Lim, J.W.; Ho, S.-H.; Show, P.L. A review on

microalgae cultivation and harvesting, and their biomass extraction processing using ionic

liquids. Bioengineered 2020, 11, 116-129.

12. Muradov, N.; Taha, M.; Miranda, A.F.; Wrede, D.; Kadali, K.; Gujar, A.; Stevenson, T.; Ball,

A.S.; Mouradov, A. Fungal-assisted algal flocculation: application in wastewater treatment

and biofuel production. Biotechnol Biofuels 2015, 8, 24.

13. Bracharz, F.; Helmdach, D.; Aschenbrenner, I.; Funck, N.; Wibberg, D.; Winkler, A.; Bohnen,

F.; Kalinowski, J.; Mehlmer, N.; Brück, T.B. Harvest of the oleaginous microalgae

Scenedesmus obtusiusculus by flocculation from culture based on natural water sources.

Front Bioeng Biotechnol 2018, 6, 200.

14. Pugazhendhi, A.; Shobana, S.; Bakonyi, P.; Nemestóthy, N.; Xia, A.; Kumar, G. A review on

chemical mechanism of microalgae flocculation via polymers. Biotechnol Rep 2019, 21, e00302.

15. Bosma, R.; van Spronsen, W.A.; Tramper, J.; Wijffels, R.H. Ultrasound, a new separation

technique to harvest microalgae. J Appl Phycol 2003, 15, 143-153.

16. Brennan, L.; Owende, P. Biofuels from microalgae—a review of technologies for production,

processing, and extractions of biofuels and co-products. Renew Sust Energy Rev 2010, 14, 557-

577.

17. Velan, M.; Saravanane, R. Pollution Abatement and Utilisation of Flue Gas for Bioenergy

Production–A Review. Intl J Emerg Technol Advanced Eng 2013, 3, 94-99.

18. Alkarawi, M.A.; Caldwell, G.S.; Lee, J.G. Continuous harvesting of microalgae biomass using

foam flotation. Algal Res 2018, 36, 125-138.

19. Najjar, Y.S.; Abu-Shamleh, A. Harvesting of microalgae by centrifugation for biodiesel

production: A review. Algal Res 2020, 51, 102046.

20. Grima, E.M.; Belarbi, E.-H.; Fernández, F.A.; Medina, A.R.; Chisti, Y. Recovery of microalgal

biomass and metabolites: process options and economics. Biotechnol Adv 2003, 20, 491-515.

21. Heasman, M.; Diemar, J.; O'connor, W.; Sushames, T.; Foulkes, L. Development of extended

shelf‐life microalgae concentrate diets harvested by centrifugation for bivalve molluscs–a

summary. Aquac Res 2000, 31, 637-659.

22. Show, K.-Y.; Lee, D.-J.; Chang, J.-S. Algal biomass dehydration. Bioresour Technol 2013, 135,

720-729.

23. Gultom, S.O.; Hu, B. Review of microalgae harvesting via co-pelletization with filamentous

fungus. Energies 2013, 6, 5921-5939.

24. de Farias Neves, F.; Demarco, M.; Tribuzi, G. Drying and Quality of Microalgal Powders for

Human Alimentation. In Microalgae-From Physiology to Application, Vítová, M., Ed.

IntechOpen: 2019; 10.5772/intechopen.89324.

25. Mujumdar, A.S. Classification and selection of industrial dryers. In Mujumdar’s Practical

Guide to Industrial Drying: Principles, Equipment and New Developments. Brossard, Canada:

Exergex Corporation, S, D., Ed. Exergex Corporation: Montreal, Canada, 2000; pp. 23-77.

26. Nappa, M.; Teir, S.; Sorsamäki, L.; Karinen, P. Energy requirements of microalgae biomass

production; CCSP Deliverable D606 Espoo 2016: Finland, Helsinki, 2016.

27. Oyinloye, T.M.; Yoon, W.B. Effect of freeze-drying on quality and grinding process of food

produce: A review. Processes 2020, 8, 354.

28. Ventura, S.; Nobre, B.; Ertekin, F.; Hayes, M.; Garciá-Vaquero, M.; Vieira, F.; Koc, M.;

Gouveia, L.; Aires-Barros, M.; Palavra, A. Extraction of value-added compounds from

microalgae. In Microalgae-Based Biofuels and Bioproducts, Elsevier: 2017; pp. 461-483.

29. Günerken, E.; D'Hondt, E.; Eppink, M.; Garcia-Gonzalez, L.; Elst, K.; Wijffels, R.H. Cell

disruption for microalgae biorefineries. Biotechnol Adv 2015, 33, 243-260.

30. Velazquez-Lucio, J.; Rodríguez-Jasso, R.M.; Colla, L.M.; Sáenz-Galindo, A.; Cervantes-

Cisneros, D.E.; Aguilar, C.N.; Fernandes, B.D.; Ruiz, H.A. Microalgal biomass pretreatment

for bioethanol production: a review. Biofuel Res J 2018, 5, 780-791.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 39: Review Isolation of Industrial Important Bioactive

39

31. Cha, K.H.; Koo, S.Y.; Lee, D.-U. Antiproliferative effects of carotenoids extracted from

Chlorella ellipsoidea and Chlorella vulgaris on human colon cancer cells. J Agr Food Chem

2008, 56, 10521-10526.

32. Pasquet, V.; Morisset, P.; Ihammouine, S.; Chepied, A.; Aumailley, L.; Berard, J.-B.; Serive, B.;

Kaas, R.; Lanneluc, I.; Thiery, V. Antiproliferative activity of violaxanthin isolated from

bioguided fractionation of Dunaliella tertiolecta extracts. Mari Drugs 2011, 9, 819-831.

33. Neumann, U.; Derwenskus, F.; Flaiz Flister, V.; Schmid-Staiger, U.; Hirth, T.; Bischoff, S.C.

Fucoxanthin, a carotenoid derived from Phaeodactylum tricornutum exerts antiproliferative

and antioxidant activities in vitro. Antioxidants 2019, 8, 183.

34. Prabakaran, G.; Sampathkumar, P.; Kavisri, M.; Moovendhan, M. Extraction and

characterization of phycocyanin from Spirulina platensis and evaluation of its anticancer,

antidiabetic and antiinflammatory effect. Int J Biol Macromol 2020, 153, 256-263.

35. Liu, Y.; Xu, L.; Cheng, N.; Lin, L.; Zhang, C. Inhibitory effect of phycocyanin from Spirulina

platensis on the growth of human leukemia K562 cells. J Appl Phycol 2000, 12, 125-130.

36. Hao, S.; Yan, Y.; Li, S.; Zhao, L.; Zhang, C.; Liu, L.; Wang, C. The in vitro anti-tumor activity

of phycocyanin against non-small cell lung cancer cells. Mar Drugs 2018, 16, 178.

37. Hao, S.; Li, S.; Wang, J.; Yan, Y.; Ai, X.; Zhang, J.; Ren, Y.; Wu, T.; Liu, L.; Wang, C.

Phycocyanin Exerts Anti-Proliferative Effects through Down-Regulating TIRAP/NF-κB

Activity in Human Non-Small Cell Lung Cancer Cells. Cells 2019, 8, 588.

38. Deniz, I.; Ozen, M.O.; Yesil-Celiktas, O. Supercritical fluid extraction of phycocyanin and

investigation of cytotoxicity on human lung cancer cells. J Supercrit Fluids 2016, 108, 13-18.

39. Gantar, M.; Dhandayuthapani, S.; Rathinavelu, A. Phycocyanin induces apoptosis and

enhances the effect of topotecan on prostate cell line LNCaP. J Med Food 2012, 15, 1091-1095.

40. Safaei, M.; Maleki, H.; Soleimanpour, H.; Norouzy, A.; Zahiri, H.; Vali, H.; Noghabi, K.

Development of a novel method for the purification of C-phycocyanin pigment from a local

cyanobacterial strain Limnothrix sp. NS01 and evaluation of its anticancer properties. Sci Rep-

UK 2019, 9, 9474-9474.

41. Harari, A.; Abecassis, R.; Relevi, N.; Levi, Z.; Ben-Amotz, A.; Kamari, Y.; Harats, D.; Shaish,

A. Prevention of atherosclerosis progression by 9-cis-β-carotene rich alga Dunaliella in apoE-

deficient mice. BioMed Res Int 2013, 2013.

42. Rao, A.; Briskey, D.; Nalley, J.O.; Ganuza, E. Omega-3 Eicosapentaenoic Acid (EPA) Rich

Extract from the Microalga Nannochloropsis Decreases Cholesterol in Healthy Individuals:

A Double-Blind, Randomized, Placebo-Controlled, Three-Month Supplementation Study.

Nutrients 2020, 12, 1869.

43. Boyd, M.R.; Gustafson, K.R.; McMahon, J.B.; Shoemaker, R.H.; O'Keefe, B.R.; Mori, T.;

Gulakowski, R.J.; Wu, L.; Rivera, M.I.; Laurencot, C.M. Discovery of cyanovirin-N, a novel

human immunodeficiency virus-inactivating protein that binds viral surface envelope

glycoprotein gp120: potential applications to microbicide development. Antimicrob Agents Ch

1997, 41, 1521-1530.

44. Bokesch, H.R.; O'Keefe, B.R.; McKee, T.C.; Pannell, L.K.; Patterson, G.M.; Gardella, R.S.;

Sowder, R.C.; Turpin, J.; Watson, K.; Buckheit, R.W. A potent novel anti-HIV protein from

the cultured cyanobacterium Scytonema varium. Biochemistry 2003, 42, 2578-2584.

45. Zainuddin, E.N.; Mentel, R.; Wray, V.; Jansen, R.; Nimtz, M.; Lalk, M.; Mundt, S. Cyclic

depsipeptides, ichthyopeptins A and B, from Microcystis ichthyoblabe. J Nat Prod 2007, 70,

1084-1088.

46. Hayashi, K.; Hayashi, T.; Kojima, I. A natural sulfated polysaccharide, calcium spirulan,

isolated from Spirulina platensis: in vitro and ex vivo evaluation of anti-herpes simplex virus

and anti-human immunodeficiency virus activities. AIDS Res Hum Retrov 1996, 12, 1463-1471.

47. Huleihel, M.; Ishanu, V.; Tal, J.; Arad, S.M. Antiviral effect of red microalgal polysaccharides

on Herpes simplex and Varicella zoster viruses. J Appl Phycol 2001, 13, 127-134.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 40: Review Isolation of Industrial Important Bioactive

40

48. Talero, E.; García-Mauriño, S.; Ávila-Román, J.; Rodríguez-Luna, A.; Alcaide, A.; Motilva, V.

Bioactive compounds isolated from microalgae in chronic inflammation and cancer. Mar

Drugs 2015, 13, 6152-6209.

49. Rumin, J.; Nicolau, E.; Junior, R.G.D.O.; Fuentes-Grünewald, C.; Picot, L. Analysis of

Scientific Research Driving Microalgae Market Opportunities in Europe. Marine drugs 2020,

18, 264, doi:10.3390/md18050264.

50. Harvey, P.J.; Ben-Amotz, A. Towards a sustainable Dunaliella salina microalgal biorefinery

for 9-cis β-carotene production. Algal Research 2020, 50, 102002,

doi:https://doi.org/10.1016/j.algal.2020.102002.

51. Gilroy, D.J.; Kauffman, K.W.; Hall, R.A.; Huang, X.; Chu, F.S. Assessing potential health risks

from microcystin toxins in blue-green algae dietary supplements. Environ Health Perspect

2000, 108, 435-439, doi:10.1289/ehp.00108435.

52. Martínez Andrade, K.A.; Lauritano, C.; Romano, G.; Ianora, A. Marine microalgae with anti-

cancer properties. Mar Drugs 2018, 16, 165.

53. Park, H.J.; Lee, Y.J.; Ryu, H.K.; Kim, M.H.; Chung, H.W.; Kim, W.Y. A randomized double-

blind, placebo-controlled study to establish the effects of spirulina in elderly Koreans. Ann

Nutr Metab 2008, 52, 322-328.

54. Dawczynski, C.; Dittrich, M.; Neumann, T.; Goetze, K.; Welzel, A.; Oelzner, P.; Völker, S.;

Schaible, A.; Troisi, F.; Thomas, L. Docosahexaenoic acid in the treatment of rheumatoid

arthritis: A double-blind, placebo-controlled, randomized cross-over study with microalgae

vs. sunflower oil. Clin Nutr 2018, 37, 494-504.

55. FEBICO. ApoX Surface Anti-Viral Spray. Availabe online:

https://www.febico.com/en/product/Spray.html (accessed on 8th October 2020).

56. Novoveská, L.; Ross, M.E.; Stanley, M.S.; Pradelles, R.; Wasiolek, V.; Sassi, J.-F. Microalgal

Carotenoids: A Review of Production, Current Markets, Regulations, and Future Direction.

Marine drugs 2019, 17, 640, doi:10.3390/md17110640.

57. Prabakaran, G.; Sampathkumar, P.; Kavisri, M.; Moovendhan, M. Extraction and

characterization of phycocyanin from Spirulina platensis and evaluation of its anticancer,

antidiabetic and antiinflammatory effect. International Journal of Biological Macromolecules

2020, 153, 256-263, doi:https://doi.org/10.1016/j.ijbiomac.2020.03.009.

58. Liu, Y.; Xu, L.; Cheng, N.; Lin, L.; Zhang, C. Inhibitory effect of phycocyanin from Spirulina

platensis on the growth of human leukemia K562 cells. Journal of Applied Phycology 2000, 12,

125-130, doi:10.1023/A:1008132210772.

59. Raposo, M.F.d.J.; de Morais, A.M.M.B.; de Morais, R.M.S.C. Carotenoids from Marine

Microalgae: A Valuable Natural Source for the Prevention of Chronic Diseases. Marine drugs

2015, 13, 5128-5155, doi:10.3390/md13085128.

60. Raposo, M.F.d.J.; de Morais, A.M.M.B.; de Morais, R.M.S.C. Influence of sulphate on the

composition and antibacterial and antiviral properties of the exopolysaccharide from

Porphyridium cruentum. Life Sciences 2014, 101, 56-63,

doi:https://doi.org/10.1016/j.lfs.2014.02.013.

61. Winwood, R.J. Recent developments in the commercial production of DHA and EPA rich oils

from micro-algae. OCL 2013, 20, D604.

62. Gustafson, K.R.; Sowder, R.C.; Henderson, L.E.; Cardellina, J.H.; McMahon, J.B.; Rajamani,

U.; Pannell, L.K.; Boyd, M.R. Isolation, Primary Sequence Determination, and Disulfide Bond

Structure of Cyanovirin-N, an Anti-HIV(Human Immunodeficiency Virus) Protein from the

CyanobacteriumNostoc ellipsosporum. Biochemical and Biophysical Research Communications

1997, 238, 223-228, doi:https://doi.org/10.1006/bbrc.1997.7203.

63. Huskens, D.; Schols, D. Algal lectins as potential HIV microbicide candidates. Mar drugs 2012,

10, 1476-1497.

64. Zainuddin, E.N.; Mentel, R.; Wray, V.; Jansen, R.; Nimtz, M.; Lalk, M.; Mundt, S. Cyclic

Depsipeptides, Ichthyopeptins A and B, from Microcystis ichthyoblabe. Journal of Natural

Products 2007, 70, 1084-1088, doi:10.1021/np060303s.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 41: Review Isolation of Industrial Important Bioactive

41

65. Watanabe, F. Vitamin B12 sources and bioavailability. Exp Biol Med 2007, 232, 1266-1274.

66. Bishop, W.; Zubeck, H. Evaluation of microalgae for use as nutraceuticals and nutritional

supplements. J Nutr Food Sci 2012, 2, 1-6.

67. Bioeconomy BW. Microalgae can produce more than just fuel. Availabe online:

https://www.biooekonomie-bw.de/en/articles/news/microalgae-can-produce-more-than-

just-fuel (accessed on 10th September 2020).

68. Pulz, O.; Gross, W. Valuable products from biotechnology of microalgae. Appl Microbiol

Biotechnol 2004, 65, 635-648.

69. Kumudha, A.; Selvakumar, S.; Dilshad, P.; Vaidyanathan, G.; Thakur, M.S.; Sarada, R.

Methylcobalamin–A form of vitamin B12 identified and characterised in Chlorella vulgaris.

Food Chem 2015, 170, 316-320.

70. Molino, A.; Rimauro, J.; Casella, P.; Cerbone, A.; Larocca, V.; Chianese, S.; Karatza, D.;

Mehariya, S.; Ferraro, A.; Hristoforou, E. Extraction of astaxanthin from microalga

Haematococcus pluvialis in red phase by using generally recognised as safe solvents and

accelerated extraction. J Biotechnol 2018, 283, 51-61.

71. Denery, J.R.; Dragull, K.; Tang, C.; Li, Q.X. Pressurized fluid extraction of carotenoids from

Haematococcus pluvialis and Dunaliella salina and kavalactones from Piper methysticum. Anal

Chim Acta 2004, 501, 175-181.

72. Haque, F.; Dutta, A.; Thimmanagari, M.; Chiang, Y.W. Intensified green production of

astaxanthin from Haematococcus pluvialis. Food Bioprod Process 2016, 99, 1-11.

73. Olaizola, M. Commercial production of astaxanthin from Haematococcus pluvialis using

25,000-liter outdoor photobioreactors. J Appl Phycol 2000, 12, 499-506.

74. Dufossé, L.; Galaup, P.; Yaron, A.; Arad, S.M.; Blanc, P.; Murthy, K.N.C.; Ravishankar, G.A.

Microorganisms and microalgae as sources of pigments for food use: a scientific oddity or an

industrial reality? Trends Food Sci Tech 2005, 16, 389-406.

75. Mendes, A.; Reis, A.; Vasconcelos, R.; Guerra, P.; da Silva, T.L. Crypthecodinium cohnii with

emphasis on DHA production: a review. J Appl Phycol 2009, 21, 199-214.

76. Kagan, M.L.; Levy, A.; Leikin-Frenkel, A. Comparative study of tissue deposition of omega-

3 fatty acids from polar-lipid rich oil of the microalgae Nannochloropsis oculata with krill oil

in rats. Food Funct 2015, 6, 185-191.

77. Kagan, M.L.; West, A.L.; Zante, C.; Calder, P.C. Acute appearance of fatty acids in human

plasma–a comparative study between polar-lipid rich oil from the microalgae

Nannochloropsis oculata and krill oil in healthy young males. Lipids Health Dis 2013, 12, 102.

78. Li, S.; Ji, L.; Shi, Q.; Wu, H.; Fan, J. Advances in the production of bioactive substances from

marine unicellular microalgae Porphyridium spp. Bioresour Technol 2019, 292, 122048.

79. Marcati, A.; Ursu, A.V.; Laroche, C.; Soanen, N.; Marchal, L.; Jubeau, S.; Djelveh, G.; Michaud,

P. Extraction and fractionation of polysaccharides and B-phycoerythrin from the microalga

Porphyridium cruentum by membrane technology. Algal Res 2014, 5, 258-263.

80. Sathasivam, R.; Radhakrishnan, R.; Hashem, A.; Abd_Allah, E.F. Microalgae metabolites: A

rich source for food and medicine. Saudi J Biol Sci 2019, 26, 709-722.

81. Ahuja, K.; Singh, S. Algae Protein Market Size By Source Global Market Insights, Inc.: USA,

2020.

82. Capelli, B.; Cysewski, G.R. Potential health benefits of spirulina microalgae. Nutrafoods 2010,

9, 19-26.

83. Group, C. Tavelmout; A Japanese algae-based protein start-up; raised total 1.7 Billion

Japanese yen. Preparing for world wide increasing protein demand. Availabe online:

https://chitose-bio.com/news/792/ (accessed on 10th September 2020).

84. Siva Kiran, R.; Madhu, G.; Satyanarayana, S. Spirulina in combating protein energy

malnutrition (PEM) and protein energy wasting (PEW)-A review. J Nutr Res 2015, 3, 62-79.

85. Ak, B.; Avsaroglu, E.; Isik, O.; Özyurt, G.; Kafkas, E.; Etyemez, M. Nutritional and

physicochemical characteristics of bread enriched with microalgae Spirulina platensis. Int J

Eng Res Appl 2016, 6, 30-38.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 42: Review Isolation of Industrial Important Bioactive

42

86. Niccolai, A.; Venturi, M.; Galli, V.; Pini, N.; Rodolfi, L.; Biondi, N.; D’Ottavio, M.; Batista,

A.P.; Raymundo, A.; Granchi, L. Development of new microalgae-based sourdough

“crostini”: Functional effects of Arthrospira platensis (spirulina) addition. Sci Rep-UK 2019, 9,

1-12.

87. Grahl, S.; Strack, M.; Weinrich, R.; Mörlein, D. Consumer-oriented product development: the

conceptualization of novel food products based on spirulina (arthrospira platensis) and

resulting consumer expectations. J Food Quality 2018, 2018.

88. Stanic-Vucinic, D.; Minic, S.; Nikolic, M.R.; Velickovic, T.C. Spirulina Phycobiliproteins as

Food Components and Complements. In Microalgal Biotechnology, IntechOpen: 2018.

89. Chiong, T.; Acquah, C.; Lau, S.; Khor, E.; Danquah, M. Microalgal-Based Protein By-Products:

Extraction, Purification, and Applications. In Protein Byproducts, Elsevier: 2016; pp. 213-234.

90. Ursu, A.-V.; Marcati, A.; Sayd, T.; Sante-Lhoutellier, V.; Djelveh, G.; Michaud, P. Extraction,

fractionation and functional properties of proteins from the microalgae Chlorella vulgaris.

Bioresour Technol 2014, 157, 134-139.

91. Chia, S.R.; Chew, K.W.; Mohd Zaid, H.F.; Chu, D.-T.; Tao, Y.; Show, P.L. Microalgal protein

extraction from Chlorella vulgaris FSP-E using triphasic partitioning technique with

sonication. Front Bioeng Biotechnol 2019, 7, 396.

92. FAO/WHO. Energy and protein requirement. Report of a Joint FAO/WHO ad hoc Expert Committee;

Geneva, 1973.

93. Becker, E.W. Micro-algae as a source of protein. Biotechnol Adv 2007, 25, 207-210.

94. Kent, M.; Welladsen, H.M.; Mangott, A.; Li, Y. Nutritional evaluation of Australian

microalgae as potential human health supplements. PloS one 2015, 10, e0118985.

95. Seghiri, R.; Kharbach, M.; Essamri, A. Functional composition, nutritional properties, and

biological activities of Moroccan Spirulina microalga. J Food Quality 2019, 2019.

96. Tibbetts, S.M.; Milley, J.E.; Lall, S.P. Chemical composition and nutritional properties of

freshwater and marine microalgal biomass cultured in photobioreactors. J Appl Phycol 2015,

27, 1109-1119.

97. Annapurna, V.; Deosthale, Y.; Bamji, M.S. Spirulina as a source of vitamin A. Plant Food Hum

Nutr 1991, 41, 125-134.

98. Annapurna, V.; Shah, N.; Bhaskaram, P.; Bamji, M.S.; Reddy, V. Bioavailability of spirulina

carotenes in preschool children. J Clin Biochem Nutr 1991, 10, 145-151.

99. Watanabe, F.; Katsura, H.; Takenaka, S.; Fujita, T.; Abe, K.; Tamura, Y.; Nakatsuka, T.;

Nakano, Y. Pseudovitamin B12 is the predominant cobamide of an algal health food,

spirulina tablets. J Agric Food Chem 1999, 47, 4736-4741.

100. Madhubalaji, C.K.; Rashmi, V.; Chauhan, V.S.; Shylaja, M.D.; Sarada, R. Improvement of

vitamin B12 status with Spirulina supplementation in Wistar rats validated through functional

and circulatory markers. J Food Biochem 2019, 43, e13038.

101. Puyfoulhoux, G.; Rouanet, J.-M.; Besançon, P.; Baroux, B.; Baccou, J.-C.; Caporiccio, B. Iron

availability from iron-fortified spirulina by an in vitro digestion/Caco-2 cell culture model. J

Agric Food Chem 2001, 49, 1625-1629.

102. Silva, J.; Alves, C.; Pinteus, S.; Reboleira, J.; Pedrosa, R.; Bernardino, S. Chapter 3.10 -

Chlorella. In Nonvitamin and Nonmineral Nutritional Supplements, Nabavi, S.M., Silva, A.S.,

Eds. Academic Press: 2019; https://doi.org/10.1016/B978-0-12-812491-8.00026-6pp. 187-193.

103. Merchant, R.E.; Phillips, T.W.; Udani, J. Nutritional supplementation with Chlorella

pyrenoidosa lowers serum methylmalonic acid in vegans and vegetarians with a suspected

vitamin B12 deficiency. J Med Food 2015, 18, 1357-1362.

104. Nakano, S.; Takekoshi, H.; Nakano, M. Chlorella pyrenoidosa supplementation reduces the

risk of anemia, proteinuria and edema in pregnant women. Plant Foods Hum Nutr 2010, 65,

25-30.

105. Chronopoulou, L.; Dal Bosco, C.; Di Caprio, F.; Prosini, L.; Gentili, A.; Pagnanelli, F.; Palocci,

C. Extraction of carotenoids and fat-soluble vitamins from Tetradesmus Obliquus microalgae:

an optimized approach by using supercritical CO2. Molecules 2019, 24, 2581.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 43: Review Isolation of Industrial Important Bioactive

43

106. Kumudha, A.; Sarada, R. Effect of different extraction methods on vitamin B12 from blue

green algae, Spirulina platensis. Pharm Anal Acta 2015, 6.

107. Tomdio, A.; Ritchie, M.; Miller, A.C. Omega-3 Fatty Acids and Cardiovascular Disease

Prevention. Am Fam Physician 2019, 100, 209-210.

108. Shahidi, F. Omega-3 oils: sources, applications, and health effects. In Marine nutraceuticals and

functional foods, Shahidi, F., Barrow, C., Eds. Taylor & Francis Inc: Bosa Roca, United States,

2008; pp. 23-61.

109. Sivaramakrishnan, R.; Incharoensakdi, A. Enhancement of total lipid yield by nitrogen,

carbon, and iron supplementation in isolated microalgae. J Phycol 2017, 53, 855-868.

110. Goncalves, E.C.; Wilkie, A.C.; Kirst, M.; Rathinasabapathi, B. Metabolic regulation of

triacylglycerol accumulation in the green algae: identification of potential targets for

engineering to improve oil yield. Plant Biotechnol J 2016, 14, 1649-1660.

111. Wang, X.; Fosse, H.K.; Li, K.; Chauton, M.S.; Vadstein, O.; Reitan, K.I. Influence of nitrogen

limitation on lipid accumulation and EPA and DHA content in four marine microalgae for

possible use in aquafeed. Front Mar Sci 2019, 6, 95.

112. Hoffmann, M.; Marxen, K.; Schulz, R.; Vanselow, K.H. TFA and EPA productivities of

Nannochloropsis salina influenced by temperature and nitrate stimuli in turbidostatic

controlled experiments. Mar drugs 2010, 8, 2526-2545.

113. Van Wagenen, J.; Miller, T.W.; Hobbs, S.; Hook, P.; Crowe, B.; Huesemann, M. Effects of light

and temperature on fatty acid production in Nannochloropsis salina. Energies 2012, 5, 731-

740.

114. Sato, N.; Tsuzuki, M.; Kawaguchi, A. Glycerolipid synthesis in Chlorella kessleri 11 h: II. Effect

of the CO2 concentration during growth. Biochim. Biophys. Acta,-Mol. Cell. Biol. Lipids 2003,

1633, 35-42.

115. Tatsuzawa, H.; Takizawa, E.; Wada, M.; Yamamoto, Y. Fatty acid and lipid composition of

the acidophilic green alga Chlamydomonas sp. 1. J Phycol 1996, 32, 598-601.

116. Kapoore, R.V.; Butler, T.O.; Pandhal, J.; Vaidyanathan, S. Microwave-assisted extraction for

microalgae: from biofuels to biorefinery. Biology 2018, 7, 18.

117. Lorente, E.; Hapońska, M.; Clavero, E.; Torras, C.; Salvadó, J. Steam explosion and vibrating

membrane filtration to improve the processing cost of microalgae cell disruption and

fractionation. Processes 2018, 6, 28.

118. Cuellar-Bermudez, S.P.; Aguilar-Hernandez, I.; Cardenas-Chavez, D.L.; Ornelas-Soto, N.;

Romero-Ogawa, M.A.; Parra-Saldivar, R. Extraction and purification of high-value

metabolites from microalgae: essential lipids, astaxanthin and phycobiliproteins. Microb

Biotechnol 2015, 8, 190-209, doi:10.1111/1751-7915.12167.

119. Martins, D.A.; Custódio, L.; Barreira, L.; Pereira, H.; Ben-Hamadou, R.; Varela, J.; Abu-Salah,

K.M. Alternative sources of n-3 long-chain polyunsaturated fatty acids in marine microalgae.

Mar drugs 2013, 11, 2259-2281.

120. Coward, T.; Fuentes-Grünewald, C.; Silkina, A.; Oatley-Radcliffe, D.L.; Llewellyn, G.; Lovitt,

R.W. Utilising light-emitting diodes of specific narrow wavelengths for the optimization and

co-production of multiple high-value compounds in Porphyridium purpureum. Bioresour

Technol 2016, 221, 607-615.

121. Gao, B.; Chen, A.; Zhang, W.; Li, A.; Zhang, C. Co-production of lipids, eicosapentaenoic

acid, fucoxanthin, and chrysolaminarin by Phaeodactylum tricornutum cultured in a flat-plate

photobioreactor under varying nitrogen conditions. J Ocean U China 2017, 16, 916-924.

122. Okafor, S.N.; Obonga, W.; Ezeokonkwo, M.A.; Nurudeen, J.; Orovwigho, U.; Ahiabuike, J.

Assessment of the health implications of synthetic and natural food colourants—A critical

review. UK J Pharm Biosci 2016, 4, 01-11.

123. Schultz, H. NutraIngredients USA: Chinese suppliers angling to snare bigger share of natural

astaxanthin market. Availabe online: https://www.nutraingredients-

usa.com/Article/2015/04/13/Chinese-suppliers-angling-to-snare-bigger-share-of-natural-

astaxanthin-market (accessed on 5th September 2020).

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 44: Review Isolation of Industrial Important Bioactive

44

124. Chacón‐Lee, T. and González‐Mariño, G. (2010), Microalgae for “Healthy” Foods—

Possibilities and Challenges. Compr Rev Food Sci Food Saf, 9: 655-

675. https://doi.org/10.1111/j.1541-4337.2010.00132.x

125. Ambati, R.R.; Phang, S.-M.; Ravi, S.; Aswathanarayana, R.G. Astaxanthin: sources, extraction,

stability, biological activities and its commercial applications—a review. Mar Drugs 2014, 12,

128-152.

126. Evans, D.A.; Rabie, M. Algal and algal extract dietary supplement composition. Google

Patents: 2009.

127. Silva, S.C.; Ferreira, I.C.; Dias, M.M.; Barreiro, M.F. Microalgae-Derived Pigments: A 10-Year

Bibliometric Review and Industry and Market Trend Analysis. Molecules 2020, 25, 3406.

128. Dasgupta, C.N. Algae as a source of phycocyanin and other industrially important pigments.

In Algal biorefinery: An integrated approach, Springer: 2015; pp. 253-276.

129. Chen, C.-Y.; Kao, P.-C.; Tan, C.H.; Show, P.L.; Cheah, W.Y.; Lee, W.-L.; Ling, T.C.; Chang, J.-

S. Using an innovative pH-stat CO2 feeding strategy to enhance cell growth and C-

phycocyanin production from Spirulina platensis. Biochem Eng J 2016, 112, 78-85.

130. Bryant, D.A.; Guglielmi, G.; de Marsac, N.T.; Castets, A.-M.; Cohen-Bazire, G. The structure

of cyanobacterial phycobilisomes: a model. Arch Microbiol 1979, 123, 113-127.

131. Galetović, A.; Seura, F.; Gallardo, V.; Graves, R.; Cortés, J.; Valdivia, C.; Núñez, J.; Tapia, C.;

Neira, I.; Sanzana, S. Use of Phycobiliproteins from Atacama Cyanobacteria as Food

Colorants in a Dairy Beverage Prototype. Foods 2020, 9, 244.

132. USFDA. Summary of Color Additives for Use in the United States in Foods, Drugs,

Cosmetics, and Medical Devices. Availabe online: https://www.fda.gov/industry/color-

additive-inventories/summary-color-additives-use-united-states-foods-drugs-cosmetics-

and-medical-devices (accessed on 17th September 2020).

133. Hejazi, M.A.; Wijffels, R.H. Milking of microalgae. Trends Biotechnol 2004, 22, 189-194.

134. Zhu, Y.-H.; Jiang, J.-G. Continuous cultivation of Dunaliella salina in photobioreactor for the

production of β-carotene. Eur Food Res Technol 2008, 227, 953-959.

135.Pourkarimi, S.; Hallajisani, A.; Alizadehdakhel, A.; Golzary, A. Factors affecting production of

beta-carotene from Dunaliella salina microalgae. Biocatal Agric Biotechnol 2020, 101771.

136. Xu, Y.; Harvey, P.J. Carotenoid production by Dunaliella salina under red light. Antioxidants

2019, 8, 123.

137. Ruane, M. Extraction of caroteniferous materials from algae. Australian patent 1977.

138. Pirwitz, K.; Flassig, R.J.; Rihko-Struckmann, L.K.; Sundmacher, K. Energy and operating cost

assessment of competing harvesting methods for D. salina in a β-carotene production

process. Algal Res 2015, 12, 161-169.

139. Singh, D.P.; Khattar, J.S.; Rajput, A.; Chaudhary, R.; Singh, R. High production of carotenoids

by the green microalga Asterarcys quadricellulare PUMCC 5.1. 1 under optimized culture

conditions. PloS one 2019, 14, e0221930.

140. Kaushal, S.; Singh, Y.; Khattar, J.; Singh, D. Phycobiliprotein production by a novel cold

desert cyanobacterium Nodularia sphaerocarpa PUPCCC 420.1. J Appl Phycol 2017, 29, 1819-

1827.

141. Katiyar, S.; Elmets, C.A.; Katiyar, S.K. Green tea and skin cancer: photoimmunology,

angiogenesis and DNA repair. The J Nutr Biochem 2007, 18, 287-296.

142. Pérez-Sánchez, A.; Barrajón-Catalán, E.; Herranz-López, M.; Micol, V. Nutraceuticals for skin

care: A comprehensive review of human clinical studies. Nutrients 2018, 10, 403.

143. Pangestuti, R.; Suryaningtyas, I.T.; Siahaan, E.A.; Kim, S.-K. Cosmetics and Cosmeceutical

Applications of Microalgae Pigments. In Pigments from Microalgae Handbook, Springer: 2020;

pp. 611-633.

144. Sawant, S.S.; Mane, V.K. Correlating the Anti-Aging Activity with the Bioactive Profile of

Chlorella emersonii KJ725233; its Toxicological Studies for a Potential use in Cosmeceuticals.

Pharmacogn Comm 2017, 7.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 45: Review Isolation of Industrial Important Bioactive

45

145. Mourelle, M.L.; Gómez, C.P.; Legido, J.L. The potential use of marine microalgae and

cyanobacteria in cosmetics and thalassotherapy. Cosmetics 2017, 4, 46.

146. Stramarkou, M.; Papadaki, S.; Kyriakopoulou, K.; Krokida, M. Recovery of functional

pigments from four different species of microalgae. IOSR J Environ Sci, Toxicol Food Technol

2016, 10, 26-30.

147. Bilal, M.; Rasheed, T.; Ahmed, I.; Iqbal, H.M. High-value compounds from microalgae with

industrial exploitability-A review. Front Biosci (Scholar edition) 2017, 9, 319-342.

148. de Jesus Raposo, M.F.; de Morais, R.M.S.C.; de Morais, A.M.M.B. Health applications of

bioactive compounds from marine microalgae. Life Sci 2013, 93, 479-486.

149. Sathasivam, R.; Ki, J.-S. A review of the biological activities of microalgal carotenoids and

their potential use in healthcare and cosmetic industries. Mar Drugs 2018, 16, 26.

150. Mäki‐Arvela, P.; Hachemi, I.; Murzin, D.Y. Comparative study of the extraction methods for

recovery of carotenoids from algae: extraction kinetics and effect of different extraction

parameters. J Chem Technol Biotechnol 2014, 89, 1607-1626.

151. Bhalamurugan, G.L.; Valerie, O.; Mark, L.; Bhalamurugan, G.L.; Valerie, O.; Mark, L.

Valuable bioproducts obtained from microalgal biomass and their commercial applications:

A review. Environ Eng Res 2018, 23, 229-241.

152. Lephart, E.D. Skin aging and oxidative stress: Equol’s anti-aging effects via biochemical and

molecular mechanisms. Ageing Res Rev 2016, 31, 36-54.

153. Davinelli, S.; Nielsen, M.E.; Scapagnini, G. Astaxanthin in skin health, repair, and disease: A

comprehensive review. Nutrients 2018, 10, 522.

154. Ryu, B.; Himaya, S.; Kim, S.-K. Applications of microalgae-derived active ingredients as

cosmeceuticals. In Handbook of marine microalgae, Elsevier: 2015; pp. 309-316.

155. Kim, H.M.; Jung, J.H.; Kim, J.Y.; Heo, J.; Cho, D.H.; Kim, H.S.; An, S.; An, I.S.; Bae, S. The

Protective Effect of Violaxanthin from Nannochloropsis oceanica against Ultraviolet B‐Induced

Damage in Normal Human Dermal Fibroblasts. Photochem Photobiol 2019, 95, 595-604.

156. Matsui, M.; Tanaka, K.; Higashiguchi, N.; Okawa, H.; Yamada, Y.; Tanaka, K.; Taira, S.;

Aoyama, T.; Takanishi, M.; Natsume, C. Protective and therapeutic effects of fucoxanthin

against sunburn caused by UV irradiation. J Pharmacol Sci 2016, 132, 55-64.

157. Begum, H.; Yusoff, F.M.; Banerjee, S.; Khatoon, H.; Shariff, M. Availability and utilisation of

pigments from microalgae. Crit Rev Food Sci Nutr 2016, 56, 2209-2222.

158. Chakdar, H.; Pabbi, S. Algal pigments for human health and cosmeceuticals. In Algal Green

Chem, Elsevier: 2017; pp. 171-188.

159. Biba, E. Protection: the sunscreen pill. Nature 2014, 515, S124-S125.

160. Godin, B.; Touitou, E. Transdermal skin delivery: predictions for humans from in vivo, ex

vivo and animal models. Adv Drug Deliver Rev 2007, 59, 1152-1161.

161. Shen, C.-T.; Chen, P.-Y.; Wu, J.-J.; Lee, T.-M.; Hsu, S.-L.; Chang, C.-M.J.; Young, C.-C.; Shieh,

C.-J. Purification of algal anti-tyrosinase zeaxanthin from Nannochloropsis oculata using

supercritical anti-solvent precipitation. J Supercrit Fluids 2011, 55, 955-962.

162. Rao, A.R.; Sindhuja, H.; Dharmesh, S.M.; Sankar, K.U.; Sarada, R.; Ravishankar, G.A.

Effective inhibition of skin cancer, tyrosinase, and antioxidative properties by astaxanthin

and astaxanthin esters from the green alga Haematococcus pluvialis. J Agric Food Chem 2013, 61,

3842-3851.

163. Bonté, F. Skin moisturization mechanisms: new data. In Proceedings of Annales

pharmaceutiques francaises; pp. 135-141.

164. Wang, H.-M.D.; Chen, C.-C.; Huynh, P.; Chang, J.-S. Exploring the potential of using algae in

cosmetics. Bioresour Technol 2015, 184, 355-362.

165. Zanella, L.; Alam, M.A. Extracts and Bioactives from Microalgae (Sensu Stricto):

Opportunities and Challenges for a New Generation of Cosmetics. In Microalgae Biotechnology

for Food, Health and High Value Products, Springer: 2020; pp. 295-349.

166. Zmijewski, M.A.; Slominski, A.T. Neuroendocrinology of the skin: An overview and selective

analysis. Dermatoendocrinol 2011, 3, 3-10.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 46: Review Isolation of Industrial Important Bioactive

46

167. Alexopoulos, A.; Chrousos, G.P. Stress-related skin disorders. Rev Endocr Metab Dis 2016, 17,

295-304.

168. Truzzi, F.; Marconi, A.; Pincelli, C. Neurotrophins in healthy and diseased skin.

Dermatoendocrinol 2011, 3, 32-36.

169. Borroni, R.; Truzzi, F.; Pincelli, C. The skin neurotrophic network in health and disease. Actas

Dermo-sifiliogr 2009, 100, 70-74.

170. Grewe, M.; Vogelsang, K.; Ruzicka, T.; Stege, H.; Krutmann, J. Neurotrophin-4 production

by human epidermal keratinocytes: increased expression in atopic dermatitis. J Invest

Dermatol 2000, 114, 1108-1112.

171. Pavlovic, S.; Daniltchenko, M.; Tobin, D.J.; Hagen, E.; Hunt, S.P.; Klapp, B.F.; Arck, P.C.;

Peters, E.M. Further exploring the brain–skin connection: stress worsens dermatitis via

substance P-dependent neurogenic inflammation in mice. J Invest Dermatol 2008, 128, 434-

446.

172. Kinkelin, I.; Bröcker, E.-B.; Koltzenburg, M.; Carlton, S.M. Localization of ionotropic

glutamate receptors in peripheral axons of human skin. Neurosci Lett 2000, 283, 149-152.

173. Sharma, K.; Sharma, D.; Sharma, M.; Sharma, N.; Bidve, P.; Prajapati, N.; Kalia, K.; Tiwari, V.

Astaxanthin ameliorates behavioral and biochemical alterations in in-vitro and in-vivo

model of neuropathic pain. Neurosc Lett 2018, 674, 162-170.

174. Koller, M.; Muhr, A.; Braunegg, G. Microalgae as versatile cellular factories for valued

products. Algal Res 2014, 6, 52-63.

175. Tredici, M.R.; Rodolfi, L.; Biondi, N.; Bassi, N.; Sampietro, G. Techno-economic analysis of

microalgal biomass production in a 1-ha Green Wall Panel (GWP®) plant. Algal Res 2016, 19,

253-263.

176. Fernandez, F.G.A.; Sevilla, J.M.F.; Grima, E.M. Microalgae: The basis of mankind

sustainability. In Case study of innovative projects-successful real cases, Llamas, B., Ed. Intech

Open: Rijeka, 2017; pp. 123-140.

177. News, A.W. Josie Maran Argan Beta Retinoid Pink Algae Serum is unlike anything else.

Availabe online: https://news.algaeworld.org/2020/01/josie-maran-argan-beta-retinoid-pink-

algae-serum-is-unlike-anything-else/#more-41347 (accessed on 18th September 2020).

178. Borowitzka, M.A. Microalgae as sources of pharmaceuticals and other biologically active

compounds. J Appl Phycol 1995, 7, 3-15.

179. Morvan, P.; Vallee, R. Effects of Chlorella extract on skin. Personal Care 2007, pp 57-64.

180. Guillerme, J.-B.; Couteau, C.; Coiffard, L. Applications for marine resources in cosmetics.

Cosmetics 2017, 4, 35.

181. Maiz, D. The underwater world: A source of inexhaustible inspiration. Parf Cosmet Actual

2007, 194, 136 - 160.

182. Campiche, R.; Sandau, P.; Kurth, E.; Massironi, M.; Imfeld, D.; Schuetz, R. Protective effects

of an extract of the freshwater microalga Scenedesmus rubescens on UV‐irradiated skin cells.

Int J Cosmetic Sci 2018, 40, 187-192.

183. Stolz, P.; Obermayer, B. Manufacturing microalgae for skin care. Cosmet Toiletries 2005, 120,

99-106.

184. Joshi, S.; Kumari, R.; Upasani, V.N. Applications of algae in cosmetics: An overview. Int J

Innov Res Sci Eng Technol 2018, 7, 1269.

185. Hussian, A. The role of microalgae in renewable energy production: challenges and

opportunities. In Marine ecology-biotic and abiotic interactions, Türkoğlu, M., Önal, U., Ismen,

A., Eds. Intech Open: 2018; https://doi: 10.5772/intechopen.73573pp. 257-283.

186. Musa, M.; Ayoko, G.A.; Ward, A.; Rösch, C.; Brown, R.J.; Rainey, T.J. Factors affecting

microalgae production for biofuels and the potentials of chemometric methods in assessing

and optimizing productivity. Cells 2019, 8, 851.

187. Bayro-Kaiser, V.; Nelson, N. Microalgal hydrogen production: prospects of an essential

technology for a clean and sustainable energy economy. Photosynth Res 2017, 133, 49-62.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 47: Review Isolation of Industrial Important Bioactive

47

188. Ganesan, R.; Manigandan, S.; Samuel, M.S.; Shanmuganathan, R.; Brindhadevi, K.; Chi,

N.T.L.; Duc, P.A.; Pugazhendhi, A. A review on prospective production of biofuel from

microalgae. Biotechnol Rep 2020, e00509.

189. Hamed, I. The evolution and versatility of microalgal biotechnology: a review. Compr Rev

Food Sci Food Saf 2016, 15, 1104-1123.

190. Cruz, Y.R.; Aranda, D.A.; Seidl, P.R.; Diaz, G.C.; Carliz, R.G.; Fortes, M.M.; da Ponte, D.; de

Paula, R.C. Cultivation systems of microalgae for the production of biofuels. In Biofuels-State

of Development, Biernat, K., Ed. Intech Open: 2018; pp. 199 - 218.

191. Abo, B.O.; Odey, E.A.; Bakayoko, M.; Kalakodio, L. Microalgae to biofuels production: a

review on cultivation, application and renewable energy. Rev Environ Health 2019, 34, 91-99.

192. USDOE. Biomass Conversion: From Feedstocks to Final Products. Availabe online:

https://www.energy.gov/sites/prod/files/2016/07/f33/conversion_factsheet.pdf (accessed on

1st October 2020).

193. Nasreen, S.; Nafees, M.; Qureshi, L.A.; Asad, M.S.; Sadiq, A.; Ali, S.D. Review of catalytic

transesterification methods for biodiesel production. In Biofuels: State of Development, Biernat,

K., Ed. Intech Open: 2018; pp. 93-119.

194. Taher, H.; Al-Zuhair, S.; Al-Marzouqi, A.H.; Haik, Y.; Farid, M.M. A review of enzymatic

transesterification of microalgal oil-based biodiesel using supercritical technology. Enzyme

Res 2011, 2011.

195. Gouveia, L.; Oliveira, A.C. Microalgae as a raw material for biofuels production. J Ind

Microbiol Biot 2009, 36, 269-274.

196. Dasan, Y.K.; Lam, M.K.; Yusup, S.; Lim, J.W.; Lee, K.T. Life cycle evaluation of microalgae

biofuels production: Effect of cultivation system on energy, carbon emission and cost balance

analysis. Sci Total Environ 2019, 688, 112-128.

197. Sheehan, J.; Camobreco, V.; Duffield, J.; Graboski, M.; Shapouri, H. An overview of biodiesel

and petroleum diesel life cycles; National Renewable Energy Lab.(NREL), Golden, CO (United

States) and US Department of Energy (DOE): 1998.

198. Wang, H.; Ji, C.; Bi, S.; Zhou, P.; Chen, L.; Liu, T. Joint production of biodiesel and bioethanol

from filamentous oleaginous microalgae Tribonema sp. Bioresour Technol 2014, 172, 169-173.

199. Khan, M.I.; Shin, J.H.; Kim, J.D. The promising future of microalgae: current status,

challenges, and optimization of a sustainable and renewable industry for biofuels, feed, and

other products. Microb Cell Fact 2018, 17, 36.

200. Rempel, A.; Machado, T.; Treichel, H.; Colla, E.; Margarites, A.C.; Colla, L.M. Saccharification

of Spirulina platensis biomass using free and immobilized amylolytic enzymes. Bioresour

Technol 2018, 263, 163-171.

201. de Farias Silva, C.E.; Bertucco, A. Bioethanol from microalgae and cyanobacteria: a review

and technological outlook. Process Biochem 2016, 51, 1833-1842.

202. Astolfi, A.L.; Rempel, A.; Cavanhi, V.A.F.; Alves, M.; Deamici, K.M.; Colla, L.M.; Costa, J.A.V.

Simultaneous saccharification and fermentation of Spirulina sp. and corn starch for the

production of bioethanol and obtaining biopeptides with high antioxidant activity. Bioresour

Technol 2020, 301, 122698.

203. Patidar, J.; Sharma, R.; Yadav, M.; Tiwari4, A. Microalgae as Sustainable Renewable Energy

feedstock for bioethanol production. Scholars Acad J Biosci (SAJB) 2017, 5, 536-542,

doi:10.21276/sajb.

204. Kim, H.M.; Oh, C.H.; Bae, H.-J. Comparison of red microalgae (Porphyridium cruentum)

culture conditions for bioethanol production. Bioresour Technol 2017, 233, 44-50.

205. Chandra, N.; Shukla, P.; Mallick, N. Role of cultural variables in augmenting carbohydrate

accumulation in the green microalga Scenedesmus acuminatus for bioethanol production.

Biocatal Agric Biotechnol 2020, 101632.

206. Harun, R.; Jason, W.; Cherrington, T.; Danquah, M.K. Exploring alkaline pre-treatment of

microalgal biomass for bioethanol production. Appl Energ 2011, 88, 3464-3467.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 48: Review Isolation of Industrial Important Bioactive

48

207. Choi, S.P.; Nguyen, M.T.; Sim, S.J. Enzymatic pretreatment of Chlamydomonas reinhardtii

biomass for ethanol production. Bioresour Technol 2010, 101, 5330-5336.

208. Harun, R.; Danquah, M.K. Influence of acid pre-treatment on microalgal biomass for

bioethanol production. Process Biochem 2011, 46, 304-309.

209. Guo, H.; Daroch, M.; Liu, L.; Qiu, G.; Geng, S.; Wang, G. Biochemical features and bioethanol

production of microalgae from coastal waters of Pearl River Delta. Bioresour Technol 2013,

127, 422-428.

210. Ho, S.-H.; Huang, S.-W.; Chen, C.-Y.; Hasunuma, T.; Kondo, A.; Chang, J.-S. Bioethanol

production using carbohydrate-rich microalgae biomass as feedstock. Bioresour Technol 2013,

135, 191-198.

211. Ho, S.-H.; Li, P.-J.; Liu, C.-C.; Chang, J.-S. Bioprocess development on microalgae-based CO2

fixation and bioethanol production using Scenedesmus obliquus CNW-N. Bioresour Technol

2013, 145, 142-149.

212. Rizza, L.S.; Smachetti, M.E.S.; Do Nascimento, M.; Salerno, G.L.; Curatti, L. Bioprospecting

for native microalgae as an alternative source of sugars for the production of bioethanol. Algal

Res 2017, 22, 140-147.

213. Wu, N.; Moreira, C.M.; Zhang, Y.; Doan, N.; Yang, S.; Phlips, E.J.; Svoronos, S.A.;

Pullammanappallil, P.C. Techno-Economic Analysis of Biogas Production from Microalgae

through Anaerobic Digestion. In Anaerobic Digestion, IntechOpen: 2019.

214. Milledge, J.J.; Nielsen, B.V.; Maneein, S.; Harvey, P.J. A brief review of anaerobic digestion of

algae for bioenergy. Energies 2019, 12, 1166-1188.

215. Jankowska, E.; Sahu, A.K.; Oleskowicz-Popiel, P. Biogas from microalgae: Review on

microalgae's cultivation, harvesting and pretreatment for anaerobic digestion. Renew Sust

Energ Rev 2017, 75, 692-709.

216. Wang, M.; Lee, E.; Dilbeck, M.P.; Liebelt, M.; Zhang, Q.; Ergas, S.J. Thermal pretreatment of

microalgae for biomethane production: experimental studies, kinetics and energy analysis. J

Chem Technol Biotechnol 2017, 92, 399-407.

217. Yu, J.; Takahashi, P. Biophotolysis-based hydrogen production by cyanobacteria and green

microalgae. In Communicating current research and educational topics and trends in applied

microbiology, Méndez-Vilas, A., Ed. FORMATEX: 2007; Vol. 1, pp. 79-89.

218. Nagarajan, D.; Lee, D.-J.; Kondo, A.; Chang, J.-S. Recent insights into biohydrogen

production by microalgae–From biophotolysis to dark fermentation. Bioresour Technol 2017,

227, 373-387.

219. Khetkorn, W.; Rastogi, R.P.; Incharoensakdi, A.; Lindblad, P.; Madamwar, D.; Pandey, A.;

Larroche, C. Microalgal hydrogen production–A review. Bioresour Technol 2017, 243, 1194-

1206.

220. Maneeruttanarungroj, C.; Lindblad, P.; Incharoensakdi, A. A newly isolated green alga,

Tetraspora sp. CU2551, from Thailand with efficient hydrogen production. Int J Hydrogen

Energ 2010, 35, 13193-13199.

221. Eroglu, E.; Melis, A. Photobiological hydrogen production: recent advances and state of the

art. Bioresour Technol 2011, 102, 8403-8413.

222. Hwang, J.-H.; Kim, H.-C.; Choi, J.-A.; Abou-Shanab, R.; Dempsey, B.A.; Regan, J.M.; Kim,

J.R.; Song, H.; Nam, I.-H.; Kim, S.-N. Photoautotrophic hydrogen production by eukaryotic

microalgae under aerobic conditions. Nature Commun 2014, 5, 1-6.

223. Torzillo, G.; Scoma, A.; Faraloni, C.; Giannelli, L. Advances in the biotechnology of hydrogen

production with the microalga Chlamydomonas reinhardtii. Crit Rev Biotechnol 2015, 35, 485-

496.

224. Radakovits, R.; Jinkerson, R.E.; Darzins, A.; Posewitz, M.C. Genetic engineering of algae for

enhanced biofuel production. Eukaryot Cell 2010, 9, 486-501.

225. Chen, W.-H.; Lin, B.-J.; Huang, M.-Y.; Chang, J.-S. Thermochemical conversion of microalgal

biomass into biofuels: a review. Bioresour Technol 2015, 184, 314-327.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 49: Review Isolation of Industrial Important Bioactive

49

226. Adnan, M.A.; Hossain, M.M. CO2 gasification of microalgae (N. Oculata)–A thermodynamic

study. In Proceedings of MATEC Web of Conferences; p. 01002.

227. USDOE. Biofuel basics. Availabe online: https://www.energy.gov/eere/bioenergy/biofuels-

basics (accessed on 1st October 2020).

228. Xu, Y.-P.; Duan, P.-G.; Wang, F.; Guan, Q.-Q. Liquid fuel generation from algal biomass via

a two-step process: effect of feedstocks. Biotechnol Biofuels 2018, 11, 83.

229. Wang, W.; Xu, Y.; Wang, X.; Zhang, B.; Tian, W.; Zhang, J. Hydrothermal liquefaction of

microalgae over transition metal supported TiO2 catalyst. Bioresour Technol 2018, 250, 474-

480.

230. Adamczyk, M.; Sajdak, M. Pyrolysis behaviours of microalgae Nannochloropsis gaditana. Waste

Biomass Valori 2018, 9, 2221-2235.

231. Bioenergy, I. Task 34: Biomass Pyrolysis Bioenergy Research Group, Aston University: United

Kingdom, 2007; pp 1-20.

232. Chye, J.T.T.; Jun, L.Y.; Yon, L.S.; Pan, S.; Danquah, M.K. Biofuel production from algal

biomass. In Bioenergy and biofuels, 1st ed.; Konur, O., Ed. CRC Press/Taylor and Francis

Group: Boca Raton, 2018; p. 621.

233. Quinn, J.C.; Davis, R. The potentials and challenges of algae based biofuels: a review of the

techno-economic, life cycle, and resource assessment modeling. Bioresour Technol 2015, 184,

444-452.

234. Leite, G.B.; Abdelaziz, A.E.; Hallenbeck, P.C. Algal biofuels: challenges and opportunities.

Bioresour Technol 2013, 145, 134-141.

235. Wen, Z. Algae for Biofuel Production. Availabe online: https://farm-

energy.extension.org/algae-for-biofuel-production (accessed on 15th September 2020).

236. Biofuelwatch. Microalgae biofuels myths and risks. Availabe online:

https://haseloff.plantsci.cam.ac.uk/resources/SynBio_reports/Microalgae-Biofuels-Myths-

and-Risks-2017.pdf (accessed on 1st October 2020).

237. Ronga, D.; Biazzi, E.; Parati, K.; Carminati, D.; Carminati, E.; Tava, A. Microalgal

biostimulants and biofertilisers in crop productions. Agronomy 2019, 9, 192-214.

238. Dineshkumar, R.; Rasheeq, A.A.; Arumugam, A.; Nambi, K.N.; Sampathkumar, P. Marine

microalgal extracts on cultivable crops as a considerable bio-fertilizer: A Review. Indian J

Tradit Know 2019, 18, 849-854.

239. Pereira, I.; Ortega, R.; Barrientos, L.; Moya, M.; Reyes, G.; Kramm, V. Development of a

biofertiliser based on filamentous nitrogen-fixing cyanobacteria for rice crops in Chile. J Appl

Phycol 2009, 21, 135-144.

240. Osman, M.E.H.; El-Sheekh, M.M.; El-Naggar, A.H.; Gheda, S.F. Effect of two species of

cyanobacteria as biofertilisers on some metabolic activities, growth, and yield of pea plant.

Biol Fert Soils 2010, 46, 861-875.

241. Chittapun, S.; Limbipichai, S.; Amnuaysin, N.; Boonkerd, R.; Charoensook, M. Effects of

using cyanobacteria and fertilizer on growth and yield of rice, Pathum Thani I: a pot

experiment. J Appl Phycol 2018, 30, 79-85.

242. Agwa, O.; Ogugbue, C.; Williams, E. Field evidence of Chlorella vulgaris potentials as a

biofertiliser for Hibiscus esculentus. Int J Agric Res 2017, 12, 181-189.

243. Garcia-Gonzalez, J.; Sommerfeld, M. Biofertiliser and biostimulant properties of the

microalga Acutodesmus dimorphus. J Appl Phycol 2016, 28, 1051-1061.

244. Dineshkumar, R.; Subramanian, J.; Gopalsamy, J.; Jayasingam, P.; Arumugam, A.;

Kannadasan, S.; Sampathkumar, P. The Impact of Using Microalgae as Biofertiliser in Maize

(Zea mays L.). Waste Biomass Valori 2019, 10, 1101-1110, doi:10.1007/s12649-017-0123-7.

245. Jochum, M.; Moncayo, L.P.; Jo, Y.-K. Microalgal cultivation for biofertilization in rice plants

using a vertical semi-closed airlift photobioreactor. PloS one 2018, 13, e0203456.

246. Dineshkumar, R.; Subramanian, J.; Arumugam, A.; Rasheeq, A.A.; Sampathkumar, P.

Exploring the microalgae biofertiliser effect on onion cultivation by field experiment. Waste

Biomass Valori 2020, 11, 77-87.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 50: Review Isolation of Industrial Important Bioactive

50

247. Bumandalai, O.; Tserennadmid, R. Effect of Chlorella vulgaris as a biofertiliser on

germination of tomato and cucumber seeds. Int J Aquat Biol 2019, 7, 95-99.

248. Nayak, M.; Swain, D.K.; Sen, R. Strategic valorization of de-oiled microalgal biomass waste

as biofertiliser for sustainable and improved agriculture of rice (Oryza sativa L.) crop. Sci Total

Environ 2019, 682, 475-484.

249. Bocchi, S.; Malgioglio, A. Azolla-Anabaena as a biofertiliser for rice paddy fields in the Po

Valley, a temperate rice area in Northern Italy. Int J Agron 2010, 2010.

250. Sholkamy, E.N.; El-Komy, H.; Al-Arfaj, A.A.; Abdel-Megeed, A.; Mostafa, A.A. Potential role

of Nostoc muscorum and Nostoc rivulare as biofertilisers for the enhancement of maize growth

under different doses of n-fertilizer. Afr J Microbiol Res 2012, 6, 7435-7448.

251. Karthikeyan, N.; Prasanna, R.; Nain, L.; Kaushik, B.D. Evaluating the potential of plant

growth promoting cyanobacteria as inoculants for wheat. Eur J Soil Biol 2007, 43, 23-30.

252. Uysal, O.; Uysal, F.O.; Ekinci, K. Evaluation of microalgae as microbial fertilizer. Eur J Sustain

Dev 2015, 4, 77-77.

253. Prasanna, R.; Kanchan, A.; Kaur, S.; Ramakrishnan, B.; Ranjan, K.; Singh, M.C.; Hasan, M.;

Saxena, A.K.; Shivay, Y.S. Chrysanthemum growth gains from beneficial microbial

interactions and fertility improvements in soil under protected cultivation. Hortic Plant J

2016, 2, 229-239.

254. Prasanna, R.; Kanchan, A.; Ramakrishnan, B.; Ranjan, K.; Venkatachalam, S.; Hossain, F.;

Shivay, Y.S.; Krishnan, P.; Nain, L. Cyanobacteria-based bioinoculants influence growth and

yields by modulating the microbial communities favourably in the rhizospheres of maize

hybrids. Eur J Soil Biol 2016, 75, 15-23.

255. Suleiman, A.K.A.; Lourenço, K.S.; Clark, C.; Luz, R.L.; da Silva, G.H.R.; Vet, L.E.M.;

Cantarella, H.; Fernandes, T.V.; Kuramae, E.E. From toilet to agriculture: Fertilization with

microalgal biomass from wastewater impacts the soil and rhizosphere active microbiomes,

greenhouse gas emissions and plant growth. Resour Conserv Recycl 2020, 161, 104924,

doi:https://doi.org/10.1016/j.resconrec.2020.104924.

256. Babu, S.; Prasanna, R.; Bidyarani, N.; Singh, R. Analysing the colonisation of inoculated

cyanobacteria in wheat plants using biochemical and molecular tools. J Appl Phycol 2015, 27,

327-338.

257. Bidyarani, N.; Prasanna, R.; Chawla, G.; Babu, S.; Singh, R. Deciphering the factors associated

with the colonization of rice plants by cyanobacteria. J Basic Microb 2015, 55, 407-419.

258. Bidyarani, N.; Prasanna, R.; Babu, S.; Hossain, F.; Saxena, A.K. Enhancement of plant growth

and yields in Chickpea (Cicer arietinum L.) through novel cyanobacterial and biofilmed

inoculants. Microbiol Res 2016, 188, 97-105.

259. Priya, H.; Prasanna, R.; Ramakrishnan, B.; Bidyarani, N.; Babu, S.; Thapa, S.; Renuka, N.

Influence of cyanobacterial inoculation on the culturable microbiome and growth of rice.

Microbiol Res 2015, 171, 78-89.

260. Prasanna, R.; Ramakrishnan, B.; Simranjit, K.; Ranjan, K.; Kanchan, A.; Hossain, F.; Nain, L.

Cyanobacterial and rhizobial inoculation modulates the plant physiological attributes and

nodule microbial communities of chickpea. Arch Microbiol 2017, 199, 1311-1323.

261. Ramakrishnan, B.; Kaur, S.; Prasanna, R.; Ranjan, K.; Kanchan, A.; Hossain, F.; Shivay, Y.S.;

Nain, L. Microbial inoculation of seeds characteristically shapes the rhizosphere microbiome

in desi and kabuli chickpea types. J Soils Sediments 2017, 17, 2040-2053.

262. Ahmed, E.; Holmström, S.J.M. Siderophores in environmental research: roles and

applications. Microb Biotechnol 2014, 7, 196-208, doi:10.1111/1751-7915.12117.

263. Årstøl, E.; Hohmann-Marriott, M.F. Cyanobacterial Siderophores—Physiology, Structure,

Biosynthesis, and Applications. Mar Drugs 2019, 17, 281.

264. Mazur, H.; Konop, A.; Synak, R. Indole-3-acetic acid in the culture medium of two axenic

green microalgae. J Appl Phycol 2001, 13, 35-42.

265. Sergeeva, E.; Liaimer, A.; Bergman, B. Evidence for production of the phytohormone indole-

3-acetic acid by cyanobacteria. Planta 2002, 215, 229-238.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 51: Review Isolation of Industrial Important Bioactive

51

266. Romanenko, E.; Kosakovskaya, I.; Romanenko, P. Phytohormones of Microalgae: Biological

Role and Involvement in the Regulation of Physiological Processes. Pt I. Auxins, Abscisic

Acid, Ethylene. Int J Algae 2015, 17, 275-289.

267. Stirk, W.; Ördög, V.; Van Staden, J.; Jäger, K. Cytokinin-and auxin-like activity in Cyanophyta

and microalgae. J Appl Phycol 2002, 14, 215-221.

268. Hussain, A.; Hasnain, S. Phytostimulation and biofertilization in wheat by cyanobacteria. J

Ind Microbiol Biot 2011, 38, 85-92.

269. Kumar, M.; Prasanna, R.; Bidyarani, N.; Babu, S.; Mishra, B.K.; Kumar, A.; Adak, A.; Jauhari,

S.; Yadav, K.; Singh, R. Evaluating the plant growth promoting ability of thermotolerant

bacteria and cyanobacteria and their interactions with seed spice crops. Sci Hortic-Amsterdam

2013, 164, 94-101.

270. Grzesik, M.; Romanowska-Duda, Z.; Kalaji, H. Effectiveness of cyanobacteria and green algae

in enhancing the photosynthetic performance and growth of willow (Salix viminalis L.)

plants under limited synthetic fertilizers application. Photosynthetica 2017, 55, 510-521.

271. Babu, S.; Bidyarani, N.; Chopra, P.; Monga, D.; Kumar, R.; Prasanna, R.; Kranthi, S.; Saxena,

A.K. Evaluating microbe-plant interactions and varietal differences for enhancing biocontrol

efficacy in root rot disease challenged cotton crop. Eur J Plant Pathol 2015, 142, 345-362.

272. Renuka, N.; Prasanna, R.; Sood, A.; Ahluwalia, A.S.; Bansal, R.; Babu, S.; Singh, R.; Shivay,

Y.S.; Nain, L. Exploring the efficacy of wastewater-grown microalgal biomass as a

biofertiliser for wheat. Environ Sci Pollut Res 2016, 23, 6608-6620.

273. Renuka, N.; Prasanna, R.; Sood, A.; Bansal, R.; Bidyarani, N.; Singh, R.; Shivay, Y.S.; Nain, L.;

Ahluwalia, A.S. Wastewater grown microalgal biomass as inoculants for improving

micronutrient availability in wheat. Rhizosphere 2017, 3, 150-159.

274. Chandini; Kumar, R.; Kumar, R.; Prakash, O. The Impact of Chemical Fertilizers on our

Environment and Ecosystem. In Research Trends in Environmental Sciences, 2019; pp. 69-86.

275. Mahapatra, D.M.; Chanakya, H.; Joshi, N.; Ramachandra, T.; Murthy, G. Algae-based

biofertilisers: a biorefinery approach. In Microorganisms for green revolution, Springer: 2018;

pp. 177-196.

276. Rojas Crespo, E.; Iglesias Hernandez, D.; Acien Fernandez, F.G.; Pozo Dengra, J. European

Patent Application: Method for obtaining concentrates of biofertilisers and biostimulants for

agricultural use from biomass of microalgae, including cyanobacteria 2020.

277. Industries, E.P. Blue Green Algae Biofertiliser. Availabe online:

https://ecologicalproduct.tradeindia.com/blue-green-algae-bio-fertilizer.html (accessed on

9th October 2020).

278. June Pharmaceutical Comp. Ltd.. Spirulina Superfertilizers: Shwe Awzar®. Availabe online:

http://www.junespirulina.com/products/Superfertilizes (accessed on 9th October 2020).

279. Biorizon Biotech. Spirulina: Algafert®. Availabe online: http://www.biorizon.es/bio-

booster/hydrolyzed/algafert/?lang=en (accessed on 9th October 2020).

280. MikroAlg Food Agric Ind Inc. Terradoc® biofertiliser. Availabe online:

http://mikroalg.com/urun-etiketi/terradoc-10 (accessed on 9th October 2020).

281. Accelergy Corp. TerraSyncTM biofertiliser. Availabe online: http://algaebiomass.org/wp-

content/gallery/2012-algae-biomass-summit/2010/06/Allen-Mark-Accelergy-Corporation-

Carbon-Capture-and-Utilisation-Challenges-and-Opportunities.pdf (accessed on 9th

October 2020).

282. MCT Tarim Ltd. Sti.. Availabe online: https://mcttarim.en.ecplaza.net/products/emek-

microbial-fertilizer_2956105 (accessed on 9th October 2020).

283. Emparan, Q.; Harun, R.; Danquah, M. Role of phycoremediation for nutrient removal from

wastewaters: A review. Appl Ecol Env Res 2019, 17, 889-915.

284. Canizares, R.; Domínguez, A. Growth of Spirulina maxima on swine waste. Bioresour Technol

1993, 45, 73-75.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 52: Review Isolation of Industrial Important Bioactive

52

285. Cañizares, R.; Rivas, L.; Montes, C.; Dominguez, A.; Travieso, L.; Benitez, F. Aerated swine-

wastewater treatment with K-carrageenan-immobilized Spirulina maxima. Bioresour Technol

1994, 47, 89-91.

286. Lim, S.-L.; Chu, W.-L.; Phang, S.-M. Use of Chlorella vulgaris for bioremediation of textile

wastewater. Bioresour Technol 2010, 101, 7314-7322.

287. Chen, X.; Li, Z.; He, N.; Zheng, Y.; Li, H.; Wang, H.; Wang, Y.; Lu, Y.; Li, Q.; Peng, Y. Nitrogen

and phosphorus removal from anaerobically digested wastewater by microalgae cultured in

a novel membrane photobioreactor. Biotechnol Biofuels 2018, 11, 190.

288. Cañizares-Villanueva, R.; Martínez-Roldán, A.; Perales-Vela, H.; Vázquez-Hernández, M.;

Melchy-Antonio, O. Bioremediation of Copper and Other Heavy Metals Using Microbial

Biomass. In Handbook of Metal-Microbe Interactions and Bioremediation, Das, S., Dash, H.R., Eds.

CRC Press: 2017; 10.1201/9781315153353-42pp. 585-602.

289. Dwivedi, S. Bioremediation of heavy metal by algae: current and future perspective. J Adv

Lab Res Biol 2012, 3, 195-199.

290. Hernández-Zamora, M.; Perales-Vela, H.V.; Flores-Ortíz, C.M.; Cañizares-Villanueva, R.O.

Physiological and biochemical responses of Chlorella vulgaris to Congo Red. Ecotox Environ

Safe 2014, 108, 72-77.

291. Saratale, R.G.; Saratale, G.D.; Chang, J.-S.; Govindwar, S.P. Bacterial decolorization and

degradation of azo dyes: a review. J Taiwan Inst Chem E 2011, 42, 138-157.

292. Bolong, N.; Ismail, A.; Salim, M.R.; Matsuura, T. A review of the effects of emerging

contaminants in wastewater and options for their removal. Desalination 2009, 239, 229-246.

293. Gavrilescu, M.; Demnerová, K.; Aamand, J.; Agathos, S.; Fava, F. Emerging pollutants in the

environment: present and future challenges in biomonitoring, ecological risks and

bioremediation. New Biotechnol 2015, 32, 147-156.

294. Tolboom, S.N.; Carrillo-Nieves, D.; de Jesús Rostro-Alanis, M.; de la Cruz Quiroz, R.; Barceló,

D.; Iqbal, H.M.; Parra-Saldivar, R. Algal-based removal strategies for hazardous

contaminants from the environment–A review. Sci Total Environ 2019, 665, 358-366.

295. Zhang, C.; Lu, J.; Wu, J.; Luo, Y. Phycoremediation of coastal waters contaminated with

bisphenol A by green tidal algae Ulva prolifera. Sci Total Environ 2019, 661, 55-62.

296. Dayana Priyadharshini, S.; Bakthavatsalam, A.K. A comparative study on growth and

degradation behavior of C. pyrenoidosa on synthetic phenol and phenolic wastewater of a coal

gasification plant. J Environ Chem Eng 2019, 7, 103079,

doi:https://doi.org/10.1016/j.jece.2019.103079.

297. Leng, L.; Wei, L.; Xiong, Q.; Xu, S.; Li, W.; Lv, S.; Lu, Q.; Wan, L.; Wen, Z.; Zhou, W. Use of

microalgae based technology for the removal of antibiotics from wastewater: A review.

Chemosphere 2020, 238, 124680.

298. Peng, F.-Q.; Ying, G.-G.; Yang, B.; Liu, S.; Lai, H.-J.; Liu, Y.-S.; Chen, Z.-F.; Zhou, G.-J.

Biotransformation of progesterone and norgestrel by two freshwater microalgae

(Scenedesmus obliquus and Chlorella pyrenoidosa): transformation kinetics and products

identification. Chemosphere 2014, 95, 581-588.

299. Jaiswal, K.K.; Kumar, V.; Vlaskin, M.; Sharma, N.; Rautela, I.; Nanda, M.; Arora, N.; Singh,

A.; Chauhan, P. Microalgae fuel cell for wastewater treatment: Recent advances and

challenges. J Water Process Eng 2020, 38, 101549.

300. Kusmayadi, A.; Leong, Y.K.; Yen, H.W.; Huang, C.Y.; Dong, C.D.; Chang, J.S. Microalgae‐

microbial fuel cell (mMFC): an integrated process for electricity generation, wastewater

treatment, CO2 sequestration and biomass production. Int J Energy Res 2020.

301. De-Bashan, L.E.; Bashan, Y. Immobilized microalgae for removing pollutants: review of

practical aspects. Bioresour Technol 2010, 101, 1611-1627.

302. Kumar, K.S.; Dahms, H.-U.; Won, E.-J.; Lee, J.-S.; Shin, K.-H. Microalgae–A promising tool

for heavy metal remediation. Ecotox Environ Safe 2015, 113, 329-352.

303. Aksu, Z.; Tezer, S. Biosorption of reactive dyes on the green alga Chlorella vulgaris. Process

Biochem 2005, 40, 1347-1361.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 53: Review Isolation of Industrial Important Bioactive

53

304. Daneshvar, N.; Ayazloo, M.; Khataee, A.; Pourhassan, M. Biological decolorization of dye

solution containing Malachite Green by microalgae Cosmarium sp. Bioresour Technol 2007, 98,

1176-1182.

305. Mona, S.; Kaushik, A.; Kaushik, C. Biosorption of reactive dye by waste biomass of Nostoc

linckia. Ecol Eng 2011, 37, 1589-1594.

306. Mohan, S.V.; Ramanaiah, S.; Sarma, P. Biosorption of direct azo dye from aqueous phase onto

Spirogyra sp. I02: Evaluation of kinetics and mechanistic aspects. Biochem Eng J 2008, 38, 61-

69.

307. Bai, X.; Acharya, K. Removal of trimethoprim, sulfamethoxazole, and triclosan by the green

alga Nannochloris sp. J Hazard Mater 2016, 315, 70-75.

308. Hom-Diaz, A.; Norvill, Z.N.; Blánquez, P.; Vicent, T.; Guieysse, B. Ciprofloxacin removal

during secondary domestic wastewater treatment in high rate algal ponds. Chemosphere 2017,

180, 33-41.

309. Villar-Navarro, E.; Baena-Nogueras, R.M.; Paniw, M.; Perales, J.A.; Lara-Martín, P.A.

Removal of pharmaceuticals in urban wastewater: High rate algae pond (HRAP) based

technologies as an alternative to activated sludge based processes. Water Res 2018, 139, 19-29.

310. Xiong, J.-Q.; Kurade, M.B.; Jeon, B.-H. Biodegradation of levofloxacin by an acclimated

freshwater microalga, Chlorella vulgaris. Chem Eng J 2017, 313, 1251-1257.

311. Xiong, J.-Q.; Kurade, M.B.; Jeon, B.-H. Can microalgae remove pharmaceutical contaminants

from water? Trends Biotechnol 2018, 36, 30-44.

312. Kong, W.; Yang, S.; Guo, B.; Wang, H.; Huo, H.; Zhang, A.; Niu, S. Growth behavior, glucose

consumption and phenol removal efficiency of Chlorella vulgaris under the synergistic effects

of glucose and phenol. Ecotox Environ Safe 2019, 186, 109762.

313. Tocher, D.R. Omega-3 long-chain polyunsaturated fatty acids and aquaculture in

perspective. Aquaculture 2015, 449, 94-107.

314. Raja, R.; Coelho, A.; Hemaiswarya, S.; Kumar, P.; Carvalho, I.S.; Alagarsamy, A. Applications

of microalgal paste and powder as food and feed: An update using text mining tool. Beni-

Suef Uni J Basic Appl Sci 2018, 7, 740-747.

315. Gill, I.; Valivety, R. Polyunsaturated fatty acids, part 1: occurrence, biological activities and

applications. Trends Biotechnol 1997, 15, 401-409.

316. Pratiwy, F.M.; Pratiwi, D.Y. The potentiality of microalgae as a source of DHA and EPA for

aquaculture feed: A review. Int J Fish Aquat Stud 8, 39 - 41.

317. Martinez-Porchas, M.; Martinez-Cordova, L.R.; Lopez-Elias, J.A.; Porchas-Cornejo, M.A.

Bioremediation of aquaculture effluents. In Microbial Biodegradation and Bioremediation,

Elsevier: 2014; pp. 539-553.

318. Amorim, M.L.; Soares, J.; Coimbra, J.S.d.R.; Leite, M.d.O.; Albino, L.F.T.; Martins, M.A.

Microalgae proteins: production, separation, isolation, quantification, and application in

food and feed. Crit Rev Food Sci Nutr 2020, 1-27.

319. Vaudel, M.; Verheggen, K.; Csordas, A.; Ræder, H.; Berven, F.S.; Martens, L.; Vizcaíno, J.A.;

Barsnes, H. Exploring the potential of public proteomics data. Proteomics 2016, 16, 214-225.

320. Fernández-Acero, F.J.; Amil-Ruiz, F.; Durán-Peña, M.J.; Carrasco, R.; Fajardo, C.; Guarnizo,

P.; Fuentes-Almagro, C.; Vallejo, R.A. Valorisation of the microalgae Nannochloropsis gaditana

biomass by proteomic approach in the context of circular economy. J Proteomics 2019, 193,

239-242.

321. Gong, Y.; Hu, H.; Gao, Y.; Xu, X.; Gao, H. Microalgae as platforms for production of

recombinant proteins and valuable compounds: progress and prospects. J Ind Microbiol Biot

2011, 38, 1879-1890.

322. Shamriz, S.; Ofoghi, H. Expression of recombinant PfCelTOS antigen in the chloroplast of

Chlamydomonas reinhardtii and its potential use in detection of malaria. Mol Biotechnol 2019,

61, 102-110.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1

Page 54: Review Isolation of Industrial Important Bioactive

54

323. Romay, C.; Gonzalez, R.; Ledon, N.; Remirez, D.; Rimbau, V. C-phycocyanin: a biliprotein

with antioxidant, anti-inflammatory and neuroprotective effects. Curr Protein Peptide Sc 2003,

4, 207-216.

324. Zheng, L.-H.; Wang, Y.-J.; Sheng, J.; Wang, F.; Zheng, Y.; Lin, X.-K.; Sun, M. Antitumor

peptides from marine organisms. Mar Drugs 2011, 9, 1840-1859.

325. Wang, X.; Zhang, X. Separation, antitumor activities, and encapsulation of polypeptide from

Chlorella pyrenoidosa. Biotechnol Progr 2013, 29, 681-687.

326. Tejano, L.A.; Peralta, J.P.; Yap, E.E.S.; Panjaitan, F.C.A.; Chang, Y.-W. Prediction of Bioactive

Peptides from Chlorella sorokiniana Proteins Using Proteomic Techniques in Combination

with Bioinformatics Analyses. Int J Mol Sci 2019, 20, 1786.

327. Dong, H.-P.; Williams, E.; Wang, D.-z.; Xie, Z.-X.; Hsia, R.-c.; Jenck, A.; Halden, R.; Li, J.;

Chen, F.; Place, A.R. Responses of Nannochloropsis oceanica IMET1 to long-term nitrogen

starvation and recovery. Plant Physiol 2013, 162, 1110-1126.

328. Tran, N.-A.T.; Padula, M.P.; Evenhuis, C.R.; Commault, A.S.; Ralph, P.J.; Tamburic, B.

Proteomic and biophysical analyses reveal a metabolic shift in nitrogen deprived

Nannochloropsis oculata. Algal Res 2016, 19, 1-11.

329. Anand, V.; Singh, P.K.; Banerjee, C.; Shukla, P. Proteomic approaches in microalgae:

perspectives and applications. 3 Biotech 2017, 7, 197.

330. Lauritano, C.; Ferrante, M.I.; Rogato, A. Marine natural products from microalgae: An-omics

overview. Mar Drugs 2019, 17, 269.

Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 25 December 2020 doi:10.20944/preprints202012.0658.v1